Next Article in Journal
Mechanisms of Photoaging and Cutaneous Photocarcinogenesis, and Photoprotective Strategies with Phytochemicals
Previous Article in Journal
Antioxidant, Biomolecule Oxidation Protective Activities of Nardostachys jatamansi DC and Its Phytochemical Analysis by RP-HPLC and GC-MS
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Silymarin as a Natural Antioxidant: An Overview of the Current Evidence and Perspectives

1
Department of Microbiology and Biochemistry, Faculty of Veterinary Medicine, Trakia University, Stara Zagora 6000, Bulgaria
2
Department of Animal Nutrition, Faculty of Agricultural and Environmental Sciences, Szent Istvan University, Gödöllo H-2103, Hungary
3
Department of Veterinary Expertise and Microbiology, Faculty of Veterinary Medicine, Sumy National Agrarian University, Sumy 40021, Ukraine
4
Odessa National Academy of Food Technology, Odessa 65039, Ukraine 
Antioxidants 2015, 4(1), 204-247; https://doi.org/10.3390/antiox4010204
Submission received: 4 January 2015 / Revised: 6 February 2015 / Accepted: 9 March 2015 / Published: 20 March 2015

Abstract

:
Silymarin (SM), an extract from the Silybum marianum (milk thistle) plant containing various flavonolignans (with silybin being the major one), has received a tremendous amount of attention over the last decade as a herbal remedy for liver treatment. In many cases, the antioxidant properties of SM are considered to be responsible for its protective actions. Possible antioxidant mechanisms of SM are evaluated in this review. (1) Direct scavenging free radicals and chelating free Fe and Cu are mainly effective in the gut. (2) Preventing free radical formation by inhibiting specific ROS-producing enzymes, or improving an integrity of mitochondria in stress conditions, are of great importance. (3) Maintaining an optimal redox balance in the cell by activating a range of antioxidant enzymes and non-enzymatic antioxidants, mainly via Nrf2 activation is probably the main driving force of antioxidant (AO)  action of SM. (4) Decreasing inflammatory responses by inhibiting NF-κB pathways is an emerging mechanism of SM protective effects in liver toxicity and various liver diseases. (5) Activating vitagenes, responsible for synthesis of protective molecules, including heat shock proteins (HSPs), thioredoxin and sirtuins and providing additional protection in stress conditions deserves more attention. (6) Affecting the microenvironment of the gut, including SM-bacteria interactions, awaits future investigations. (7) In animal nutrition and disease prevention strategy, SM alone, or in combination with other hepatho-active compounds (carnitine, betaine, vitamin B12, etc.), might have similar hepatoprotective effects as described in human nutrition.

1. Introduction

Silymarin (SM) is a C25 containing flavonoid mixture, extracted from the Silybum marianum (milk thistle) plant. Today’s standardized (according to its silibinin, often called silybin, content) SM extract contains approximately 65% to 80% flavonolignans (silybin A and silybin B, isosilybin A, isosilybin B, silychristin and silydianin), with small amounts of flavonoids, and approximately 20% to 35% of fatty acids and polyphenolic compounds possessing a range of metabolic regulatory effects [1]. Silybin was discovered as the first member of a new family of natural compounds called flavonolignans in 1959 [2] and it is known as the predominant and primary active ingredient in SM [3,4]. That is why compounds containing milk thistle ingredients showing silybin content and silybin antioxidant, as well as other activities in various model systems, are used to explain the biological activity of SM. In particular, SM has been the gold standard drug to treat liver disorders of different etiologies and milk thistle extracts have been used as traditional herbal remedies (“liver tonics”) for almost 2000 years. Therefore, SM is most well known for its antioxidant and chemoprotective effects on the liver [5,6,7,8,9,10,11] and it is often prescribed and self-prescribed as a complementary and alternative hepatoprotective medicine [12]. SM is being studied as a hepato-, neuro-, nephro- and cardio-protective ingredient due to its strong antioxidant and tissue regenerative properties [12,13,14,15,16,17]. There is a range of recent comprehensive reviews covering various routes and mechanisms of action of SM in animal models and human trials [13,14,15,16,17] very often referring to its antioxidant properties. However, it seems likely that direct antioxidant (AO) activity of polyphenols does not contribute directly to the antioxidant defence of the body [18,19] and only limited work has been carried out to explore SM/silybin impact on the induction of cellular antioxidant defence via the modulation of various transcription factors, including Nrf2 and NF-κB and respective gene and protein expressions. Potential molecular proliferative signaling targets for anti-cancer activity of silibinin include the receptor tyrosine kinase, STAT, androgen receptor and NF-κB pathways [20], however, anti-cancer activity of SM is beyond the scope of the present review. Therefore, this review focuses on evaluating recent studies on SM (silibinin) antioxidant effects in various in vitro and in vivo model systems in the context of its contribution to the antioxidant systems regulation and participation in cell signaling.

2. Absorption and Metabolism of Silibinin

SM and its main constituent silibinin sources, metabolism and bioavailability have already been reviewed extensively [3,21,22]. It has been shown that after oral consumption silibinin is characterised by comparatively low availability, e.g., in rats it is only about 0.95% [23]. In fact, after the oral administration of the standardized milk thistle extract Legalon, flavonolignans were rapidly absorbed and eliminated [22] with a half-life for silibinin of 6 h [24,25,26]. The main biotransformation route of silybin and derivatives was identified to be conjugation [3]. It is interesting to note that silibinin, similar to other flavonoids is recognized by the body as a foreign matter and quickly metabolized via phase II enzymes. Indeed, oral consumption of silibinin was associated with a significant increase in both glutathione S-transferase (GST) and quinone reductase (QR) activities in liver, lung, stomach, skin and small bowel in a dose- and time-dependent manner [27]. Silibinin present in the systemic circulation was found mainly in conjugated form [28,29]. In fact, after oral SM administration to healthy volunteers, only 10% [30] to 17% [31] of the total silibinin in the plasma was found in the free unconjugated form. Indeed, mono-, di-, and sulpho-glucuronides are shown to be formed, and 31 metabolites have been identified [32]. Indeed, silibinin in humans and rats shows fast elimination of both the free and conjugated forms with the mean elimination half-life being 6.32 h [24]. Therefore, similar to other flavonoids, after oral consumption silibinin, the main constituent of SM, is characterised by comparatively low availability, fast metabolism and its concentration in plasma is mainly in nano-molar range and only in some cases reaching micro-molar concentrations.

3. Antioxidant Systems of the Body

During evolution, living organisms have developed specific antioxidant protective mechanisms to deal with reactive oxygen species (ROS) and reactive nitrogen species (RNS) [33]. Therefore, it is only the presence of natural antioxidants in living organisms which enable them to survive in an oxygen-rich environment. The general term “antioxidant systems” describes these mechanisms, which are diverse and responsible for the protection of cells from the actions of free radicals. The protective antioxidant compounds are located in organelles, subcellular compartments, or the extracellular space enabling maximum cellular protection to occur. Thus, antioxidant systems of the living cell include three major levels of defence [33,34,35,36,37].
The first level of defence is responsible for prevention of free radical formation by removing precursors of free radicals or by inactivating catalysts and consists of three antioxidant enzymes, namely superoxide dismutase (SOD), glutathione peroxidase (GSH-Px), and catalase (CAT) plus metal-binding proteins. Since the superoxide radical is the main free radical produced in physiological conditions in the cell, SOD (EC 1.15.1.1) is considered to be the main element of the first level of antioxidant defence in the cell. This enzyme dismutates the superoxide radical in the following reaction:
2 O2+ 2 H+ → H2O2 + O2
The hydrogen peroxide formed by SOD action can be detoxified by GSH-Px or CAT which reduce it to water as follows:
H2O2 + 2 GSH → GSSG + 2 H2O
2 H2O2 → 2 H2O + O2
Transition metal ions also accelerate the decomposition of lipid hydroperoxides into cytotoxic products such as aldehydes, alkoxyl radicals, and peroxyl radicals. Therefore, metal-binding proteins (transferrin, lactoferrin, haptoglobin, hemopexin, metallothionein, ceruloplasmin, ferritin, albumin, myoglobin, etc.) also belong to the first level of defence. Unfortunately, the first level of antioxidant defence in the cell is not sufficient to completely prevent free radical formation and some radicals do escape through this level, initiating lipid peroxidation and causing damage to polyunsaturated fatty acids (PUFAs), DNA and proteins. Therefore, the second level of defence consists of chain-breaking antioxidants—vitamin E, ubiquinol, carotenoids, vitamin A, ascorbic acid, uric acid, and some other antioxidants. Glutathione (GSH) and thioredoxin systems also have a substantial role in the second level of antioxidant defence. Chain-breaking antioxidants inhibit peroxidation by keeping the chain length of the propagation reaction as small as possible. Therefore, they prevent the propagation step of lipid peroxidation by scavenging peroxyl radical intermediates in the chain reaction.
However, even the second level of antioxidant defence in the cell is not able to prevent lipid and protein oxidation and some biological molecules are damaged. In this case, the third level of defence is based on systems that eliminate damaged molecules or repair them. This level of antioxidant defence includes lipolitic (lipases), proteolytic (peptidases or proteases) and other enzymes (DNA repair enzymes, ligases, nucleases, polymerases, proteinases, phospholipases, various transferases, etc.). All these antioxidants are operating in the body in association with each other forming an integrated antioxidant system. The cooperative interactions between antioxidants in the cell are vital for maximum protection from the deleterious effects of free radicals and toxic products of their metabolism.
Therefore, the antioxidant defences include several options [33,34,35,36,37]:
  • Decrease localized oxygen concentration
  • Decrease activity of pro-oxidant enzymes and improve efficiency of electron chain in the mitochondria and decreasing electron leakage leading to superoxide production
  • Prevention of first-chain initiation by scavenging initial radicals by inducing various transcription factors (e.g., Nrf2 and NF-κB) and ARE-related synthesis of AO enzymes (SOD, GSH-Px, CAT, glutathione reductase (GR), GST, etc.)
  • Vita-gene activation and synthesis and increased expression of protective molecules (GSH, Thioredoxins, heat shock proteins (HSPs), sirtuins, etc.)
  • Binding metal ions (metal-binding proteins) and metal chelating
  • Decomposition of peroxides by converting them to non-radical, nontoxic products (Se-GSH-Px);
  • Chain breaking by scavenging intermediate radicals such as peroxyl and alkoxyl radicals (vitamins E, C, GSH, uric acid, ubiquinol, bilirubin, etc.)
  • Repair and removal of damaged molecules (methionine sulfoxide reductase (Msr), DNA-repair enzymes, chaperons, etc.)

4. Antioxidant Properties of Silymarin (SM)

It should be noted that SM can contribute to the antioxidant defences in different ways. Firstly, by direct free radical scavenging. Secondly, by preventing free radical formation by inhibiting specific enzymes responsible for free radical production, or by maintaining the integrity of electron-transport chain of mitochondria in stress conditions. Thirdly, by participating in the maintenance of optimal redox status of the cell by activating a range of antioxidant enzymes and non-enzymatic antioxidants, mainly via transcription factors, including Nrf2 and NF-κB. Finally, by activating an array of vitagenes, responsible for the synthesis of protective molecules, including HSP, thioredoxin (Trx), sirtuins, etc., and providing additional protection in stress conditions. In most studies pure silybin, as the main component of SM, was used, however, in some cases SM also showed antioxidant action in vivo.

4.1. Direct Free Radical Scavenging

The effects of silibinin on the formation of ROS and eicosanoids by human platelets, white blood and endothelial cells were studied [32]. Silibinin showed to be a strong scavenger of HOCI (IC50 7 µM), but not of O2 (IC50 > 200 μM) produced by human granulocytes. Similarly, production of O2 and NO in isolated rat Kupffer cells were inhibited by silibinin in a dose-dependent manner, with IC50 80 μM [38,39]. Indeed, silybin has no superoxide anion scavenging capability but was able to significantly decrease (at 20 μM) hem-mediated low dencity lipoprotein (LDL) oxidation and showed slight inhibition of hydroxyl radical formation [40]. The rate constants of silybin with OH radical (1.8 × 1010 dm/mol/s) is diffusion controlled, suggesting that silybin is a potent free radical scavenger [41]. Indeed, silibinin (2.5 μM) significantly decreased the concentration of H2O2 in Ab1–42-stressed neurons and prevented oxidative-related injuries [42]. Furthermore, silybin (10 μM and higher) is shown to have protective activity in ameliorating DNA damage in a model system [41]. Treatment in vitro with silibinin significantly inhibited spontaneous O2 and H2O2 release and TNF-α production by monocytes from pre-eclamptic women. The main effect of silibinin was obtained at 50 μM concentration [43]. Thus, the authors concluded that silibinin exerts anti-oxidative and anti-inflammatory effects on monocytes from pre-eclamptic pregnant women by inhibiting the in vitro endogenous release of ROS and TNF-α production.
Antioxidant action of SM (equal to 62 μM of silybin) has been shown in various model systems. For example, IC50 for H2O2 was 38 μM, while IC50 for NO was 266 μM [44]. It was shown that SM inhibited 82.7% lipid peroxidation of linoleic acid emulsion; while BHA, BHT, alpha-tocopherol and Trolox indicated inhibition of 83.3, 82.1, 68.1 and 81.3% respectively on peroxidation of linoleic acid emulsion at the same concentration [45]. In another in vitro study free radical scavenging activity and antioxidant properties of SM (>200 μM) were showed by four different assays [46]. It is important to mention that the free radical scavenging activity of pure individual compounds of the SM is reported to vary considerably, with silydianin and silychristin being 2-10-fold more active than the silibinin and on a mass basis, SM is shown to be about 8-fold more potent than silibinin as a free radical scavenger [47]. Indeed, SM and silybin (silibinin) are not the same compounds and their AO activities could differ substantially depending on their concentrations in the target tissues. Therefore, in the following sections a distinction will be made between AO activities of SM and its active constituent silybin.
It seems likely, that SM/silibinin AO effects could be on the cells further upstream of the ROS and TNFa release. In fact, SM has been shown to alter trafficking within cells [48,49] and affect the energy status of the cell [50], so the ROS scavenging effects, if any, could be secondary. Furthermore, many of the effects of SM/silibinin described in vitro occur at concentrations not currently achievable in humans or animals. Therefore, a direct scavenging ROS by silibinin in biological systems (except the gut) is not likely to substantially contribute to the antioxidant protection.

4.2. Protective Effects of Silybin on Mitochondria, a Main Source of Free Radical Production in the Cell

Mitochondria are the primary cellular consumers of oxygen and contain numerous redox enzymes capable of transferring single electrons to oxygen, generating the ROS superoxide (O2). It is well documented that mitochondrial enzymes known to generate ROS include the tricarboxylic acid (TCA) cycle enzymes aconitase and α-ketoglutarate dehydrogenase; the electron-transport chain (ETC) complexes I, II and III; pyruvate dehydrogenase and glycerol-3-phosphate dehydrogenase; dihydroorotate dehydrogenase; the monoamine oxidases (MAO) A and B; and cytochrome b5 reductase [51]. Furthermore, mitochondrial insults, including oxidative damage itself, can cause an imbalance between ROS production and removal, resulting in net ROS production. For example, ROS can induce protein modifications, lipid peroxidation and mitochondrial DNA damage, which ultimately results in mitochondrial dysfunction [52]. Many studies have focused on the detrimental effects of ROS, but it is now clear that mitochondrially generated ROS are also involved in regulating intracellular signal transduction pathways that result in cell adaptation to stress [53].
One of the mechanisms responsible for the decrease in oxidative stress is the protective effect of SM/silibinin on mitochondrial structure and function. Indeed SM protects mitochondria from pathological events by triggering pro-survival cell signaling. For example, silibinin supplementation is shown to optimize electron-transport chain, decreasing electron leakage and ROS formation and directly reducing activities of ROS-producing enzymes in the mitochondria. In rats subjected to ischemia/reperfusion (I/R), compared with the control group, a severe impairment of mitochondrial bioenergetics was observed. SM prevented the most significant changes that occurred in mitochondria during I/R (decreased ATP levels, membrane potential and state 3 respiration) and the associated cell dysfunction [54].
Silibinin (100–500 μM), was evaluated for its protective effect against beta-adrenergic agonist isoproterenol-induced injury in cultured rat neonatal cardiac myocytes [55]. It was shown that silibinin addition was associated with increased SOD activity and upregulated mitochondrial membrane potential and with a prevention of mitochondrial dysfunction and cell injury [54]. Silibinin, at a concentration as low as 10 μM, fully mitigated the rise in metabolic flow-driven ROS formation in perfused rat hepatocytes. In addition, studies on isolated liver mitochondria revealed that this low dose of silibinin depressed ROS production linked to the electron transfer chain activity [56]. It has been shown that cold preservation and warm reperfusion of the rat liver were associated with increased lipid peroxidation and superoxide anion generation, as well as with decreased GSH, mitochondrial ATP content and respiratory control ratio (RCR). However, preservation conducted in presence of silibinin (100 μM) improved parameters affected by preservation and reperfusion. Indeed, silibinin promoted an increase of ATP and RCR by 39 and 16% respectively and decreased oxidative stress to values observed in livers never preserved nor perfused [57]. It has been suggested that the uncoupler-like activity of dehydrosilybin could be the basis of its ROS modulation effect in various experimental systems. In fact, dehydrosilybin uncoupled the respiration of isolated rat heart mitochondria with a very high potency in suppressing ROS formation in isolated rat heart mitochondria with IC(50) = 0.15 μM [58]. It is interesting to note that silibinin in mitochondria was far more effective than its effect in a purely chemical system generating superoxide or in cells capable of oxidative burst, where the IC(50) for dehydrosilybin exceeds 50 μM. Changes in mitochondrial respiratory complexes in fatty hepatocytes were also attenuated by silibinin-vitamin E complex (15 mg vitamin E and 47 mg silybin) fed to rats with a major protective effect on Complex II subunit CII-30 [59]. Similarly, silybin (0.4 g/kg) in complex with phospholipid (SILIPHOS) was effective in decreasing severe oxidative stress and preserving hepatic mitochondrial bioenergetics and mitochondrial proton leak and ATP reduction in nonalcoholic steatohepatitis induced by the methionine- and choline-deficient (MCD) diet [60]. Silymarin oil (10 mL/kg/BW) significantly increased levels of membrane fluidity and membrane potential of liver mitochondria [61]. It has been suggested that protective mechanism of action of silibinin (50–200 μM) in intrastriatal MPP+-injected rats may be linked to maintenance of mitochondrial bioenergetics and integrity [62]. An in vitro study demonstrated that silibinin inhibits the activity of ROS-generating monoamine oxidase (MAO) that catalyzes the oxidative deamination of monoamines [63]. Similarly, silymarin oil (10 mL/kg/BW) decreased MAO activity in mice livers [61]. The formation of leukotrienes via the 5-lipoxygenase pathway was indicated to be strongly inhibited by silibinin. In particular, in human granulocytes IC50-values of 15 μM and 14.5 μM silibinin were detected for LTB4 and LTC4/D4/E4/F4 formation respectively. However, much higher silibinin concentrations (45–69 μM) were necessary to inhibit the cyclooxygenase pathway [38]. Rats exposed to a carcinogen (1,2-di-methylhydrazine DMH) showed increased activities of phase I enzymes (cytochrome b5, cytochrome b5 reductase, cytochromeP450, cytochromeP450 reductase, cytochromP4502E1) in the liver and colonic mucosa as compared to control rats. Silibinin supplementation (50 mg/kg/BW) modulates the xenobiotic metabolizing enzymes including decreasing activity of ROS-producing cytochrome b5 reductase [64]. Although energy generation in the mitochondrion is an essential and extremely important process for cell survival, excessive mitochondrial ROS production also has detrimental consequences for the cell and the whole body.
The different behaviour of silymarin/silibinin in normal and cancerous cells should be mentioned. In particular, SM is shown to have a protective effect against diabetes-induced cardiomyocyte apoptosis [65] as well as apoptosis caused by various toxicants (see Section 4.4), while it causes apoptosis in cancerous cells. For example, SM effectively suppressed cell growth in a dose- and time-dependent manner, and arrested cell cycle progression at G1/S phase in human ovarian cancer line A2780s and PA-1 cells via up-regulation of p53, p21, and p27 protein expression, and down-regulation of CDK2 protein expression [66]. Furthermore, in the aforementioned study the silymarin treatment significantly induced apoptosis in A2780s and PA-1 cells by increasing Bax and decreasing Bcl-2 protein expression, and activation of caspase-9 and caspase-3. Indeed, anti-cancer properties of SM are well characterised but they are beyond the scope of the review.

4.3. Inhibition of Free-Radical Producing Enzymes by Silybin

4.3.1. Xanthine Oxidase

Xanthine oxidase (XO) catalyzes the terminal two steps of purine degradation, converting hypoxanthine to uric acid and is considered to be a critical source of both O2 and H2O2 in inflammatory diseases [67]. In particular, XO in its oxidase form is considered to be a main source of oxidative stress and destructive free radicals in ischemia reperfusion injury associated with heart attacks, stroke and spinal cord injury, as well as being a destructive force in myocardial or renal hypoxia and infarctions [68]. It seems likely that silibinin can affect XO activity, and inhibition of XO by silibinin in the steady-state appears to be independent of the silibinin concentration. Furthermore, in the presence of 50 μM silibinin, the production of superoxide by XO was decreased by approximately 20% relative to the control at 100 μM xanthine [69]. When O2 production was induced in xanthine/xanthine oxidase system, silybin inhibited uric acid formation with an IC50 of 32.2 μM [68]. Indeed, inhibition in uric acid formation occurred parallel with inhibition of cytochrome C reduction (e.g., with O2 formation) suggesting that the observed effects are due to the inhibition of xanthine oxidase activity rather than the scavenging of O2 radicals.

4.3.2. NADPH Oxidase

ROS generation by NADPH oxidase enzyme complex plays a critical role in several physiological processes including host defense, posttranslational modification of proteins, cell differentiation and regulation of gene expression [70]. Silybin and its derivatives inhibited the NADPH oxidase activity in PMA-stimulated cell lysate in a concentration dependent manner. The most distinct inhibition was observed when cell lysate was pre-incubated with 5,7,4′′-trimethylsilybin possessing high lipid solubility. Indeed, 50% inhibition of NADPH oxidase activity was achieved when using 5 μM of 5,7,4′′-trimethylsilybin and 10 μM of silybin at final concentrations [68]. It is interesting to note that As-intoxicated rats showed a significant (p < 0.05) up-regulation of myocardial NADPH (NOX) oxidase sub units such as NOX2 and NOX4 as well as Keap-1, and down-regulation of Nrf2 and heme oxygenase (HO-1) protein expressions. Pre-administration of silibinin (75 mg/kg/BW) remarkably recovered all those altered parameters to near normalcy in As-induced cardiotoxic rat [71]. Co-administration of silibinin (75 mg/kg/BW) along with As-resulted in a reversal of As-induced biochemical changes in the kidney, accompanied by a significant decrease in lipid peroxidation and an increase in the level of renal antioxidant defense system [72]. It also decreased the NADPH oxidase, iNOS and NF-κB over expression by As and upregulated the Nrf2 expression in the renal tissue. The effect of silybin in cultures of mouse podocytes and in the OVE26 mouse, a model of type 1 diabetes mellitus and diabetic nephropathy was investigated. Exposure of podocytes to high glucose increased intracellular superoxide production by 60%, NADPH oxidase activity by 90%, NOX4 expression by 100%, and the number of apoptotic cells by 150%, effects that were completely blocked by 10 μM silybin. These in vitro observations were confirmed by similar in vivo findings. The kidney cortex of vehicle-treated control OVE26 mice displayed greater NOX4 expression and twice as much superoxide production than cortex of silybin-treated mice [73]. Therefore, the in vitro data demonstrated that silybin decreased superoxide generation in cultured podocytes and the in vivo study showed a similar effect in the kidney cortex. This is not surprising considering that oxidative stress and enhanced generation of ROS is the underlying mechanism of injury to multiple cell types.

4.4. Antioxidant Protective Properties of SM (Silibinin) in Prevention of Toxic Effects of Various Chemicals

The liver is a primary organ involved in biotransformation of food toxicants and drugs. Toxins absorbed from the intestinal tract first enter the liver resulting in a variety of liver disorders. Hepatic disorders are mainly caused by toxic chemicals (alcohol), xenobiotics (carbon tetrachloride, chlorinated hydrocarbons), anticancer (doxorubicin, cisplatin), immunosuppressant (cyclosporine), analgesic anti-inflammatory (paracetamol, thioacetamide), anti-tubercular (isoniazid, rifampicin) drugs, radiations (gamma radiations), heavy metals (cadmium, arsenic), mycotoxins (aflatoxin, T-2 toxin, fumonisins, ochratoxins), galactosamine, lipopolysaccharides, etc. [74]. Increasing evidence supports the important role of oxidative stress as a key mechanism of hepatotoxicity. Therefore, the aforementioned toxicants were used in various experimental models to assess antioxidant protective properties of SM. These are especially important, since SM is primarily used in clinical practice to protect, support and repair the liver.

4.4.1. Arsenic

As-induced toxicity is associated with increased production of ROS, elevation in lipid peroxidation, protein carbonylation and DNA oxidation, and disruption of the cell cycle that results in apoptosis [75]. Indeed, the toxic effect of arsenic was indicated by significantly decreased activities of AO enzymes (SOD, CAT, GSH-Px, GST, GR and glucose 6-phosphate dehydrogenase (G6PDH)) along with non-enzymatic antioxidants (GSH, vitamins C and E) in the rat liver; [76] and heart [71] and decreased Nrf2 and HO-1 activities in the heart [71]. Similar As-induced changes were seen in the rat kidney, including decreased activity of AO enzymes (SOD, CAT, GSH-Px, GSH, etc.), GSH, ascorbic acid and vitamin E, increased thiobarbituric acid reactive substances (TBARS), lipid peroxides and protein carbonyls, overexpression of NADPH oxidase, iNOS and NF-κB [72]. Silibinin was effective in counteracting all the parameters of oxidative stress. Indeed, administration of silibinin (50–75 mg/kg/BW) exhibited a significant preservation of antioxidant defences by maintaining activity of AO enzymes and non-enzymatic antioxidants in As-treated rats [76,77,78]. All the changes were supported by reduction of DNA damage in hepatocytes and histopathological observations of the liver.

4.4.2. Carbone Tetrachloride

Carbone tetrachloride (CCl4) is considered as an important hepatotoxin due to its severe oxidative effect on the liver. Indeed, metabolism of CCl4 via CYP2E1 to highly reactive ROS plays a major role in the mode of action of the toxicant [79]. It is well established that CCl4 inhibits AO enzymes (SOD, GSH-Px and CAT) and GSH in liver samples [80], increases the secretion of ALT, aspartate transaminase (AST) and ALP due to hepatic injuries caused by ROS [81] and enhances TBARS in the liver [82] and serum [83]. It has been shown that practically all elevated indexes of the oxidative stress caused by CCl4 were restored almost to the initial physiological levels by SM treatment (25–100 mg/kg/BW) [80,81,82,83,84,85,86,87] and hepatic injuries were significantly decreased. It is important to mention that in a model system based on CCl4-induced liver fibrosis in mice, microarray analysis showed that SM downregulated the expression levels of cytoskeleton organization genes and mitochondrion electron-transfer chain genes, such as cytochrome c oxidase, Cox6a2, Cox7a1, and Cox8b genes [88]. Furthermore, SM (0.1–1.0 g/kg diet) had a protective effect against CCl4-induced liver damage and AO system defences in common carp [89].

4.4.3. Mycotoxins

It has been shown that major mycotoxins (aflatoxins, AF; ochratoxin A, OTA; T-2 toxin and fumonisins, FM) cause oxidative stress and apoptosis in the cell [90,91,92]. Indeed, mycotoxins generate ROS, which induce lipid and protein oxidation leading to changes in membrane integrity, cellular redox signaling, and in the antioxidant status of the cells. Silibinin is shown to be a potent protective compound against apoptosis and cytotoxicity caused by OTA [93]. In fact, silibinin in doses from 130 to 260 μM prevented chromatin condensation, caspase-3 activation, apoptotic DNA fragmentation and lipid peroxidation that were induced by OTA, H2O2 and ActD/TNF-α, respectively. Furthermore, silibinin (0.04–26.0 μM) had hepatoprotective effects against OTA- or LPS-mediated TNF-alpha release and also reduced the cytotoxicity of both compounds [94]. Pre-treatment of the rats with SM (100 mg/kg/BW) prior to AFB1 was found to show significant protection against AFB1-induced liver damage, as evidenced by a significant lowering of the activity of the serum enzymes glutamic-oxaloacetic transaminase (GOT) and glutamic-pyruvic transaminase (GPT) and enhanced hepatic reduced GSH status [95]. SM (100 mg/kg/BW) also ameliorated toxic liver damage caused by FB₁ in BALB/c mice by preventing elevation of the levels of caspase-8 and tumor necrosis factor-alpha mediators [96].

4.4.4. Thioacetamide

Thiacetamide (TAA) is a potent hepatotoxicant which requires metabolic activation by the mixed-function oxidases. In the development of TAA-induced toxicity, free radicals are thought to play a critical role [97] and TAA-induced cirrhosis resulted in oxidative stress in the liver, as observed by a high level of lipid peroxidation, accompanied by distorted status of antioxidants [98]. SM is shown to have a protective effect against oxidative stress caused by TAA in rats. Liver fibrosis was induced in male Sprague Dawley rats by TAA administration (0.03% w/v) in drinking water for a period of 12 weeks. SM (50 mg/kg/BW) was shown to significantly increase hepatic antioxidant enzymes (SOD, CAT and GSH-Px) activity in the TAA-treated rats [99]. Indeed, a significant decrease in TAA-induced liver damage was observed in SM-pre-treated rats indicated by a reduction in serum GPT, GOT and alkaline phosphatase (ALP), restoration of the antioxidant system (SOD, CAT, GR, GSH-Px and GST) and decreased lipid peroxidation [100]. Increases in hepatic levels of malondialdehyde (MDA) and protein carbonyls in rats associated with TAA administration were partially blocked by SM consumption [101]. Furthermore, GSH depletion and heat shock protein-47 gene expressions were also decreased in response to SM administration.

4.4.5. Cisplatin

Cisplatin (CDDP) is a chemotherapeutic drug widely used against a variety of cancers and its nephrotoxicity is mainly due to ROS production and oxidative stress [102]. It was shown that CDDP caused decreased activities of AO enzymes (SOD and GSH-Px) and GSH, increased MDA in rat liver [103,104] and significantly elevated serum activities of lactate dehydrogenase (LDH) and creatine kinase (CK) [104]. SM (100 mg/kg/BW) significantly prevented the cisplatin-evoked disturbances in the above-mentioned antioxidant indexes [103,104]. Furthermore, in vitro pre-treatment with 25–200 μM of SM significantly protected against cisplatin-induced cell death in a dose-dependent manner [105], inhibited apoptotic cascade and increased cell viability in the HEI-OC1 cells [106].

4.4.6. Mn Toxicity

It is well known that manganese toxicity in animals is associated with increased oxidative stress, apoptosis and inflammation [107]. In fact, animals exposed to manganese chloride were characterised by a significant increase in TBARS levels associated with a decrease of enzymatic (SOD, CAT, GSH-Px) and non-enzymatic (GSH, Vitamin C) antioxidants [108,109], increased lipid and protein oxidation [110], DNA fragmentation and urinary hydrogen peroxide [111]. Co-administration of SM (100 mg/kg/BW) to Mn-treated rats significantly restored antioxidant defences and attenuated oxidative damages observed in the liver, kidney and brain [108,109,110,111].

4.4.7. Benzo[a]pyrene

In experiments with benzo(a)pyrene (B(a)P), an environmental contaminant causing oxidative stress, the protective effect of SM/silibinin (400–500 μM) was shown. It was associated with decreasing DNA damage and apoptosis [112], preventing protein thiol oxidation and AO enzyme (SOD, GSH-Px, CAT, GR, GST) inhibition in the hemolysate [113], restoring redox status, modulating glutathione metabolizing enzymes, decreasing formation of protein oxidation products and changing the level of antioxidant enzymes (5 mM; [114]). SM (50–250 mg/kg/BW) showed substantial protective effects against B(a)P-induced damages by inhibiting phase I detoxification enzyme CYP1A1 and modulating phase II conjugating enzymes (glutathione-S-transferase, epoxide hydroxylases, uridine diphosphate glucuronosyltransferases, NAD(P)H: quinone oxidoreductase 1, sulfotransferases) in rat liver [115].

4.4.8. Doxorubicin

Doxorubicin (DOX, also called adriamycin) is an effective chemotherapeutic agent.However, its side effect is cardiotoxicity, caused by oxidative stress, mitochondrial dysfunction and apoptosis [116]. It has been shown that DOX-induced biochemical and histopathological alterations in rat testis could be prevented and/or protected by SM (50 mg/kg/BW). The SM protective and preventive effects which result in the reduction of DOX-induced carbonyl stress and DNA damage [117], decrease in the plasma creatine phosphokinase (CPK), LDH, creatinine and urea, myocardial MDA contents and renal tissue contents of MDA and GSH have also been shown [118]. Similarly, SM (16 mg/kg/BW) significantly decreased an oxidative stress in DOX-treated rats by decreasing MDA and DNA fragmentation in the liver [119]. A protective influence of SM (60 mg/kg/BW) on the heart and liver tissue against toxicity induced by DOX in rats was associated with a prevention of an increase in AST and CK serum activity [120].

4.4.9. Ethanol

Oxidative stress plays an important role in the pathogenesis of alcoholic liver damage. During ethanol metabolism, ROS and RNS are formed causing oxidative stress [121]. Supplementation with a standardized SM (200 mg/kg/BW) attenuated an oxidative stress caused by acute ethanol administration, including prevention of elevation of serum ALT activity, decrease in hepatic GSH and enhanced lipid peroxidation, and increased hepatic TNF production [122]. Similarly, SM (100–200 mg/kg/BW) effectively protected liver from alcohol-induced oxidative stress as evidenced by reducing ALT and AST activities in serum, increasing SOD and GSH-Px activities and decreasing MDA content in the liver [123]. Furthermore silybin (250 mg/kg/BW) normalized alcohol-altered AO parameters in the whole blood hemolysate of mice, including AO enzymes (SOD, CAT, GR, GSH-Px), GSH and TBARS [124].

4.4.10. Other Toxicants

The protective effects of similar doses of SM/silybin on the antioxidant systems of the body are also shown in pyrogallol toxicity [125,126,127], bleomicyn-injected mice [128], acrolein toxicity in mice [129], diethylnitrosamine-induced oxidative stress in rat liver [130], sodium fluoride [131] and sodium nitrite [132] toxicity.
Indeed, in animal experiments and in vitro studies SM/silibinin blocked the oxidative toxicities of As, CCl4, various mycotoxins, thioacetamide, cisplatin, Mn overload, benzo[a]pyrene, doxorubicin, ethanol and some other toxicants by decreasing an abnormally high production of ROS in living cells and restoring signal transduction and gene expression. In the aforementioned studies the free radical scavenging and antioxidant properties of SM and silibinin are demonstrated by: (a) restoration of the endogenous AO enzymes (SOD, CAT, GSH-Px, GR and GST) and non-enzymatic antioxidants (vitamins E and C) in the liver and other tissues of stressed animals; (b) increased intracellular concentration of GSH in liver and other tissues; (c) decreased lipid and protein oxidation, detected as reduced MDA/TBARS and carbonyl content; (d) decreased DNA fragmentation/damage and apoptosis; (e) reduced secretion of ALT, AST, ALP from the liver into the plasma due to hepatic injuries caused by ROS; (f) restored Nrf2 and HO-1 activities; (g) reduced NF-κB expression and concentration of pro-inflammatory cytokines, including tumor necrosis factor. It seems likely that the use of SM before or soon after a toxic insult is more protective than in chronic liver disease [133]. A number of in vitro and in vivo studies have also shown an anticancer protective effect of SM and silibinin in various in vitro and in vivo model systems and they were comprehensively reviewed [124,125,126,127,128,129,130,131,132,133,134,135,136,137,138]. However, because of limitation of space in this review they will not be included in the analysis.

4.5. Is Effective Silybin Concentration in Vitro Achievable in Vivo?

As mentioned above, because of their poor absorption and rapid elimination, plasma values of silybin will rarely exceed micromole concentrations. Therefore, the SM concentration used in most of the aforementioned studies are not achievable in the in vivo target tissues, where the silybin concentration is usually at least 10-100-fold lower. Indeed, the active free silibinin concentration in plasma after oral consumption of SM, depending on dose of supplementation, could be in the range 0.2–2.0 μM. For example, in healthy volunteers, after an oral administration of SM (equivalent to 120 mg silibinin), total (unconjugated + conjugated) silibinin concentration in plasma was 1.1–1.3 μg/mL [139]. Lower bioavailability of silibinin was shown in another study with higher dose of SM consumption (equivalent to 240 mg silibinin), where maximum silibinin concentration was about 0.7 μM [24]. Comparable silibinin concentrations (0.3–1.9 μM) were reported in healthy volunteers receiving about 240 mg silibinin [140]. More recently, silybin A maximal concentration in human plasma was shown to be 0.2–0.6 μM and silybin B concentrations comprised about 30% of those shown for silybin A [22]. The concentrations of free and conjugated silibinin in the target tissues of mice have been shown to reach their maximum levels within 1 h after 50 mg/kg/BW silibinin administration. The Cmax of free/conjugated silibinin in the tissues were 18.3/11.8; 8.9/5.8; 255/560; 2.9/8.9; 5.3/12.6; 12.0/21.9 μM for liver, lung, stomach, skin, prostate and pancreas, respectively [27]. Indeed, the average concentration of most plant polyphenols in plasma rarely exceeds 1–2 μM in healthy subjects [141,142,143]. Furthermore, the effect of SM in physiologically relevant concentrations would be relatively limited when taking into account concentrations of other antioxidants in plasma. For example, vitamin C (26.1 to 84.6 μM), vitamin E (20–30 μM), urate (150–200 μM), and albumin (several hundred μM) are found in plasma [19,144] and polyphenol contribution to plasma TAC will certainly be less than 2% [19]. A similar situation could be found in animals (chickens) where antioxidant concentrations in plasma are in the same range—vitamin C: 61–66 μM; [145,146,147]; vitamin E: 13–15 μM [34]; and total antioxidant value in plasma could vary from 336.9 μM [147] up to 740–830 μM, [148]. Indeed, the general concentration of antioxidants in human plasma is about 1000 μM [149], therefore 20–50 μM additional antioxidant from dietary sources would be required to make a significant contribution to systemic antioxidant capacity [150]. While SM displays potent antioxidant activity in vitro, the bioactive forms of flavonoids in vivo are not those forms found in plants (i.e., flavonoid glycosides) due to their extensive biotransformation in the small intestine and hepatic metabolism upon absorption. Therefore, metabolic modifications occurring in vivo may substantially influence the antioxidant activity of dietary flavonoids. Silibinin, similar to other flavonoids, is recognized by the body as foreign matter and quickly metabolized via phase II enzymes.
The aforementioned data clearly indicates that SM protects cells against the damaging effects of oxidative/nitrosative stress and inflammation. It also prevents unwanted cell proliferation in cancer cells [66] as well as pathogen infection [151].

5. Oxidative Stress and Transcription Factors

Oxidation-reduction (redox) based regulation of gene expression appears to be a fundamental regulatory mechanism in cell biology. Indeed, it is well established that the level of ROS in the body is rigidly controlled by the antioxidant systems and an excess of free radical production and severe oxidative stress often leads to widespread oxidative damage and cell death. However, a basal level of oxidative stress is essential for cell survival. In fact, a moderate level of oxidative stress, induced by a variety of stressors, can create adaptive responses and improve adaptive ability to stressful challenges [152]. Therefore, a concept of the cellular antioxidant defence has recently been revised with special attention paid to cell signaling. Indeed, in animals redox-signaling pathways use ROS to transfer signals from different sources to the nucleus to regulate a number of various functions including growth, differentiation, proliferation and apoptosis. Various transcription factors are involved in a regulation of the antioxidant defence system [153,154,155,156]. These pathways operate in a coordinated manner and several are critically important for animals to cope with oxidative stress insults. They include Keap1/Nrf2, NF-κB, Mapk and AP-1 [157]. In particular great attention has been paid to a basic leucine zipper transcription factor, Nuclear factor-erythroid-2- (NF-E2-) related factor 2 (Nrf2).

5.1. Transcription Factor Nrf2

The recent information on Nrf2 functions can be summarized as follows. Based on the existing evidence, Nrf2 is considered to be the redox-sensitive master regulator of oxidative stress signaling. Growing evidence has demonstrated that the Nrf2 antioxidant response pathway plays an important role in the cellular antioxidant defense by activating a wide variety of genes involved in early defence reactions of higher organisms [158,159]. Indeed, Nrf2 is considered to have a significant role in adaptive responses to oxidative stress, being responsible for the induction of the expression of various antioxidants to combat oxidative and electrophilic stress [160]. In particular, critical components of the cellular antioxidant defense mechanisms include the ROS scavengers, phase II detoxification enzymes, and other detoxification proteins, which contain antioxidant response elements (AREs) in their promoter regions [161,162,163]. There is considerable experimental evidence suggesting that under normal physiological conditions, Nrf2 is kept in the cytoplasm by forming an inactive complex with the negative regulator, Kelch-like-ECH-associated protein 1 (Keap1), which is anchored to the actin cytoskeleton. In fact, cytosolic protein Keap1 sequesters Nrf2 in the cytoplasm and forwards it to a Cul3-based E3 ligase with the following rapid ubiquitin-proteasome degradation resulting in a short (about 20 min) half-life of Nrf2 under normal conditions (for review see [164]. Therefore, Keap1 acts as a redox sensor and upon exposure to oxidative or electrophilic stress, critical cysteine thiols of Keap1 are modified and Keap1 loses its ability to ubiquitinate Nrf2. Furthermore, phosphorylation of Nrf2 at specific serine and/or tyrosine residues also causes an Nfr2-Keap1 dissociation resulting in Nrf2 release and translocation to the nucleus. In the nucleus, Nrf2 combines with a small musculoaponeurotic fibrosarcoma protein called Maf to form a heterodimer [165]. Indeed, Nrf2 binds to the ARE in the upstream promoter region of genes encoding various antioxidant molecules, leading to the expression of antioxidant proteins, thiol molecules and other protective molecules. This includes enzymes of the first line of the antioxidant defence, namely SOD, GSH-Px and CAT, as well as detoxification enzymes (HO-1, NQO1, GST), GSH-related proteins (γ-GCS), NADPH-producing enzymes and others stress-response proteins contributing to counteracting oxidative and inflammatory damage [166,167]. The aforementioned proteins are vital to the maintenance and responsiveness of a cell antioxidant system. Therefore, an orchestrated change in gene expression via Nrf2 and the ARE is responsible for a synergistic protective effect against oxidative stress [168]. There is considerable experimental evidence suggesting that Nrf2 is controlled through a complex transcriptional/epigenetic and post-translational network that ensures its activity increases in response to redox disturbances, inflammation, growth factor stimulation and nutrient/energy fluxes, orchestrating adaptive responses to diverse forms of stress [169]. Nrf2 can be activated by various mechanisms, including stabilization of Nrf2 via Keap1 cysteine thiol modification and phosphorylation of Nrf2 by upstream kinases [170,171]. Beyond the activation of synthesis of antioxidant molecules, Nrf2 also contributes to adaptation by up-regulating the repair and degradation of damaged macromolecules, and by modulating intermediary metabolism conducting metabolic reprogramming during stress [167]. It is important to note that the expression of Nrf2 has been shown throughout human tissue, with high expression in detoxification organs, including the liver and kidney [160]. It has been suggested that low intensity oxidative stress is mainly sensed by Keap1/Nrf2 system with the following downstream up-regulation of the protective AO genes. Intermediate oxidative stress also increases the activity of antioxidant enzymes, but mainly via NF-κB and AP-1 pathways. At both low and intermediate intensity oxidative stresses, MAP-kinases and other kinases are also involved in signal sensing and cellular response, leading to enhanced antioxidant potential [157]. Emerging evidence also indicates that Nrf2 actively interacts with other transcription factors, including heat shock factor (Hsf1; [172]) providing additional options for AO system regulation.
It has been shown that naturally occurring triterpenoids affect Nrf2/ARE pathway and its downstream targets by triterpenoids were shown to be protective in different diseases [173,174,175]. Furthermore, Nrf2 activation and subsequent regulation of the transcription of the plethora of protective genes may be increased due to the chemical modifications of the triterpenoid structure [173].

5.1.1. SM and Nrf2 Regulation

An interesting approach to explain health-promoting properties of polyphenols could be to explore the modulation of signaling pathways, e.g., the Nrf2/Keap1/ARE and NF-κB pathways, resulting in increased expression of genes encoding for cytoprotective molecules, including AO enzymes and phase II detoxification enzymes. In addition, decreased expression of NF-κB-regulated genes would reduce production of pro-inflammatory cytokines. Indeed, it has been shown that some phytochemicals (sulforaphane, curcumin, EGCG, diallyl sulfide, resveratrol, etc.) at concentrations as low as 10 μM can activate Nrf2 signaling by inducing phosphorylation of Nrf2 via activation of upstream protein kinases and/or through direct interaction with Keap1 cysteine thiols [170]. It is interesting to note that anthocyanins activate the antioxidant transcription factor Nrf2, at a concentration as low as 1 μM [123]. Silymarin’s ability to upregulate the master antioxidant coordinator Nrf2 has also been investigated in various in vitro and in vivo model systems. Initially, the SM-containing supplement Protandim (ashwagandha, bacopa extract, green tea extract, SM, and curcumin) was shown to induce HO-1 in a cell culture primarily through Nrf2 activation [176]. Protandim also induces Nrf2 nuclear localization and antioxidant enzyme expression [177]. However, it is difficult to distinguish the SM effect in the mixture. Furthermore, it was shown that SM (a water soluble formulation Legalon-SIL containing a succinate-modified active compounds) at 50 and 100 μM significantly upregulated Nrf2 protein levels in CON1 cells after 48h treatment [178]. SM (104 μM) protected human cells against the damage from the herbicide paraquat via triggering anti-oxidant related genes, including Nrf2, NQO1 and HO-1 [179]. In an in vitro experiment with human hepatic HepG2 cells, SM (100 μM) was shown to be protective against AAPH-induced oxidative stress and apoptosis and the antioxidant potential of SM was correlated with induction of antioxidant genes including HO-1, NQO-1, γ-GCLC and SOD via transcriptional activation of Nrf2 [180].
In a dietary rat model of non-alcoholic steatohepatitis there was an increase in nuclear translocation of Nrf2 protein in SM-fed animals [181] which could enhance the protective effect against oxidative stress. Numerically, there was also a trend in increasing Nrf2 endogenous expression protein in rat liver tissue damaged by CCl4 by pre-treatment with SM (100 mg/kg/BW [182]) as well as in an alcohol-pyrazole-fed rat model (50 mg/kg/BW [183]). It seems likely that the effect of SM on the Nrf2 activation in various oxidative-stress related animal models depends on the dose of SM used. For example, at 10 mg/kg/BW SM showed protective effects against dimerumic acid-induced changes in GSH and HO-1 levels, but did not change Nrf2 levels [166]. Indeed, SB-treatment (75 mg/kg/BW) ameliorated arsenic-induced nephrotoxicity by abrogation of oxidative stress, inflammation and apoptosis in rats [72]. It is interesting to note that beyond the preventive action of SB, in rats fed silibinin alone there was a significant increase in Nrf2 mRNA levels in the kidney. The same SB-treated rats were characterised by increased antioxidant defences as indicated by increased renal SOD and GSH-Px activities, as well as enhanced levels of GSH, vitamins E and C. In fact, SB-induced activation of Nrf2 and HO-1 played a significant role in cellular defence against As-induced oxidative cardiac damage. In a recent study, co-treatment with SM (86 mg/kg/BW) significantly enhanced the antioxidant defence systems of the ethanol-consuming mice via the activation of Nrf2 and HO-1 expression in damaged livers [184]. Furthermore, SM (200 mg/kg/BW) significantly up-regulated HO-1 and Nrf-2 in the liver of AAPH-treated mice [180].
Therefore, it is clear from the aforementioned data that SM/silibinin, similar to other polyphenols, acts as an indirect antioxidant through its ability to induce Nrf2 transactivation, which is responsible for the synthesis of an array of antioxidant molecules decreasing oxidative stress and providing effective protection in various stress conditions. This changes an attitude to SM, a compound that has been assumed to act typically as a free radical scavenger and traditional antioxidant. It is still not clear if physiologically-relevant (up to 2 μM) silybin concentrations would upregulate Nrf2, but in vivo results obtained with various model systems clearly indicated an activation of Nrf2 by silibinin. In general, most of the experimental studies suggest that dietary SM/silibinin activates antioxidant pathways such as Nrf2/HO1 and downregulates NFκB (see Section 5.2.1. below), MMPs [44,101,185], PPAR [186,187], HIF-1 [188,189,190] and STAT [191,192,193] pathways.

5.1.2. Oxidability and Pro-Oxidant Properties of SM

It has been shown that the oxidation of Keap1 cysteine thiols can be mediated by some polyphenols. It is an interesting fact that among flavonoids, the higher their intrinsic potential to generate oxidative stress and redox cycling, the stronger their potency as inducers of ARE-mediated gene expression [194]. It was hypothesized that low concentrations of polyphenols could generate H2O2 and activates Nrf2 signaling, inducing cell adaptation to oxidative stress [195]. Thus, polyphenols act as nutritional “medicines”, which might have a preventative nature, rather than functioning as therapeutic agents. Therefore, the activation of Nrf2-ARE signaling by antioxidant polyphenols to induce various AO molecules is probably attributable to their prooxidant activity [170]. It seems likely that the same can be applied for silybin. In fact, pure silybin was found to be unstable whilst silybin in SM was stable in buffers from pH 1.0 to 7.8. The metabolism of silybin was more severe in its pure form compared to silybin in SM, as tested in a range of biological fluids including plasma, intestinal fluid and liver homogenates. It would appear that components in SM have a stabilization effect on its main component silybin [196]. Therefore, it could well be that in some in vitro experiments with cell cultures, silibinin could be oxidised and its effects would be a reflection of the oxidised form of the flavonoid or ROS generated during silybin oxidation. For example, silybin and 2,3-dehydrosilybin were shown to chelate transition metals, especially Cu2+ and pro-oxidant properties of such a complex in vitro have been shown [197]. Similarly, silybin can cause pro-oxidant effects via iron-catalyzed oxidation with the subsequent generation of reactive oxygen species [198]. Silibinin has been shown to be a strong pro-oxidative agent, i.e., it was able to oxidize NADH in vitro in the presence of peroxidase and H2O2 [50,199]. This pro-oxidative action results from the production of free-radical derivatives and subsequent NADH oxidation [200,201]. Treatment with SM (50, 100, or 200 μM) for 24 h affected the cellular redox status and induced a dose-dependent increase in ROS generation in HepG2 cells. There was also a dose-dependent decrease in intracellular GSH level and decreased total AO potential in HepG2 cells [202]. Silibinin induced cell death in human breast cancer cell lines MCF7 and MDA-MB-231 and it was attenuated by antioxidants, suggesting that the effect of silibinin was dependent on ROS generation [203]. Similarly, treatment of HT29 cells with silibinin increased the intracellular ROS level [204]. It has been suggested that SM at 100 mg/kg/BW without stressors might exert pro-oxidant effects in animals. In fact, the levels of NO and MDA increased in CCl4 and SM (100 mg/kg/BW)-treated groups, while SM at lower dose levels (25 and 50 mg/kg/BW) did not alter the NO and MDA content in the rat hippocampus. The concentration of total thiol molecules increased in the SM50 group and showed a remarkable decrease in CCl4 and SM100 groups [117]. From the aforementioned results, we can suggest that silybin can affect Nrf2 activity via its oxidation-reduction potential.

5.2. Transcription Factor NF-κB

The nuclear factor-kappa B (NF-κB) is a widely expressed, inducible transcription factor that has been implicated in regulation of many cellular processes, including inflammation. NF-κB, consisting of a group of five related proteins that are capable of binding to DNA, is activated by a wide range of stimuli including oxidative stress. It regulates the transcription of a range of genes, including pro-inflammatory cytokines and leukocyte adhesion molecules, acute phase proteins and anti-microbial peptides [205,206,207]. Similar to Nrf2, in normal physiological conditions, NF-κB is found in cytoplasm in an inactive state associated with the inhibitory IκB proteins preventing its binding to target sites. Activation of NF-κB is considered to be an effective mechanism of host defense against infection and stress [208]. Indeed, in response to stimuli, including cytokines and other stressors, IκB proteins are rapidly phosphorylated by IκB kinase on specific serine residues, with following ubiquitination, and degradation by the 26S proteasome. The resulting release of NF-κB and subsequent translocation to the nucleus orchestrates the transcription of target genes, responsible for cell survival and involved with inflammation, immunity, apoptosis, cell proliferation and differentiation [209].
There are also other stimuli implicated in NF-κB activation including cell-surface receptors, inhibitory κB kinases, IκB proteins and factors that regulate the posttranslational modification of the Rel proteins, etc. [206,207,208,209]. Current information indicates that there are complex interplay/crosstalk mechanisms between Nrf2 and NF-κB pathways. On the one hand, NF-κB pathway is inhibited by several Nrf2 activators. On the other, NF-κB may directly antagonize the transcriptional activity of Nrf2 (for review see [206]). In recent years, several compounds, including various polyphenols, have been isolated from plants that have inhibitory activities against multiple components of NF-κB activation pathway and it seems likely that SM could also have inhibitory activity on NF-κB.

5.2.1. SM and NF-κB Regulation

The anti-inflammatory activity of phenolic compounds has been demonstrated in a number of in vitro and in vivo studies and polyphenols may affect inflammation mainly as modulators of inflammatory redox signaling pathways [210]. The polyphenolic compounds express anti-inflammatory activity by modulating the expression of pro-inflammatory genes such as cyclooxygenase, lipoxygenase, nitric oxide synthases and several important cytokines, mainly acting through nuclear factor-κB and mitogen-activated protein kinase signaling [210,211]. Due to the large number of studies that have demonstrated regulatory effects of SM/silybin on the expression of NF-κB in various in vitro and in vivo models there is insufficient space in this review to analyze all of them, so we will focus only on recent investigations addressing the issue. It is well established that various plant-derived polyphenols can suppress TNF-α activated, NF-κB-associated inflammatory pathways both in vitro and in vivo. These polyphenols include SM as well as curcumin, resveratrol, genistein, epigallocatechin gallate, flavopiridol, emodin, morin isoliquiritigenin, naringenin, ellagic acid, apigenin, kaempferol, catechins, myricetin, xanthohumol, fisetin, vitexin, escin, mangostin and others [210]. SM has long been known to inhibit the activation of inflammatory NF-κB signaling [212]. Recent studies have confirmed that the protective anti-inflammatory effects of SM/silybin could be mediated by their inhibitory potential on the NF-κB, which is a key transcriptional factor for numerous genes involved in regulation of inflammation, immune system, cell differentiation, survival, apoptosis, etc. Here, we give some examples of participation of SM/silibinin in NF-κB pathway inhibition.

5.2.1.1. In Vitro Studies

Silybin is a potent inhibitor of NF-κB activation. Indeed, Manna et al. [213] tested silybin in a number of in vitro human cell experimental systems and found it inhibited TNF-mediated NF-κB activation in a dose-dependent manner with maximum inhibition at 50 μM and it regulated NF-κB 100 times more effectively than aspirin. Furthermore, silibinin (80 μM) protected glial cells against peroxide-induced ROS formation, ATP depletion, and cell damage [214]. Interestingly, the inhibition of peroxide-induced ROS by silybin was partially attenuated by inhibitors of NF-κB or protein kinase C (PKC), suggesting an involvement of NF-κB and PKC signaling pathways. Similarly, a significant decrease in NF-κB was observed, when PBMCs from pre-eclamptic patients were cultured with silibinin at 5 μM and 50 μM concentrations and stimulated with LPS [215]. Moreover, silibinin (50–200 μM) inhibited the nuclear translocation of nuclear factor NF-κB through inhibition of the phosphorylation of IκBα and suppressed NF-κB transcriptional activity in stimulated HMC-1 cells [216]. SM is also effective in inhibiting T cell activation and proliferation, by acting on pathways of NF-κB activation/translocation. In particular, in CD4+ splenocytes from C57/Bl6 mice, SM (50 μM) significantly inhibited CD4+ cells proliferation, inhibited IL-2 and IFN-gamma production and blocked nuclear translocation of transcription factor NF-κB. Moreover, SM inhibited p65/NF-κB phosphorylation in CD4+ T cell [217]. The significant upregulation of oxidative stress biomarkers including MDA, TNF-like, IFN-γ and IL-1β genes was observed as well as NF-κB, COX-2 and iNOS proteins expression occurred upon heat stress in chicken hepatocytes. Furthermore, AO enzyme (SOD, CAT, GR) activities decreased. SM (259 μM) was able to normalize the expression of all of these biomarkers in heat-induced chicken hepatocytes [218]. However, there is a need for more studies with pure silybin at physiologically-relevant concentrations (0.2–2 μM) in order to confirm its modulating effect on the NF-κB pathway.

5.2.1.2. In Vivo Studies

It has been shown that silibinin (50 mg/kg/BW) had a protective effect against d-galactose-induced senescence due to promotion of cellular oxidoreductase activities and inhibition of NF-κB activation and ROS production [219]. Histopathological and immunohistochemical studies in the kidney of rats also shows that silibinin (75 mg/kg/BW) markedly reduced the toxicity of As and preserved the normal histological architecture of the renal tissue, inhibited the caspase-3 mediated tubular cell apoptosis and decreased the NADPH oxidase, iNOS and NF-κB overexpression by As [72]. In a model of experimental ischemic stroke in male rats, silibinin (100 mg/kg) significantly downregulated NF-κB in ischemic brain tissue after stroke [189]. In an experimental nonalcoholic steatohepatitis model, mice fed a methionine-choline deficient diet consistently had increased NF-κB p65 and p50 binding activity, while silibinin administration (20 mg/kg/BW/i.p.) significantly decreased the activity of both subunits [220]. It was shown that CCl4 treatment of mice enhanced the NF-κB expression in the liver and induced hepatic fibrosis, while SM (200 mg/kg/BW) reduced the CCl4-induced hepatic NF-κB and improved CCl4-induced liver fibrosis [88]. Protective effect of silibinin (150 mg/kg/BW/intragastric) in the rat model of cerebral ischemia was associated with decreased expression of the Bax, NF-κB protein and Bax, NF-κB mRNA in the brain [221]. SM is also shown to have NF-κB inhibiting effects. For example, mRNA and protein expressions of NF-κB were significantly up-regulated in d-galactosamine induced hepatotoxic rats and treatment with SM (25 mg/kg/BW) significantly down-regulated the expressions of these genes [222]. Furthermore, it has been shown that the serum levels of TGF-β1, NF-κB, and IL-6 were significantly elevated in the samples from cirrhosis rats while SM (50 g/kg) significantly prevented the aforementioned changes [223] and SM (250 mg/kg/BW) suppressed NF-κB signaling cascade in alcoholic liver fibrosis in guinea pigs [224,225]. Overall, SM/silibinin targets multiple signaling pathways, including NF-κB, towards inhibiting the secretion of pro-inflammatory cytokines. Together these findings suggest that the inhibitory effect of SM/silibinin on NF-κB signaling in a variety of cells, as well as in in vivo studies, could be an important mechanism of its anti-inflammation efficacy, and that targeting this signaling pathway by silibinin may have potential applications in the clinic and veterinary practice.

6. Effect of SM on Vitagene Expression

The term “vitagene” was introduced in 1998 by Rattan [226] who wrote “Our survival and the physical quality of life depends upon an efficient functioning of various maintenance and repair processes. This complex network of the so-called longevity assurance processes is composed of several genes, which may be called vitagenes. Later, the vitagene concept has been further developed by Calabrese and colleagues [53,227,228,229,230,231,232,233,234,235,236,237] and major pro-survival mechanisms controlled by homodynamic vitagene network are shown in Table 1. In accordance with Calabrese et al. [227,228,229,230,231,232,233,234,235,236,237,238] the term vitagenes refers to a group of genes that are strictly involved in preserving cellular homeostasis during stress conditions and the vitagene family includes heat shock proteins (Hsps), heme oxygenase-1 (Hsp32, HO-1), Hsp60 and Hsp70, the thioredoxins (Trx)/thioredoxins reductase (TR) system and sirtuins. The list of potential candidates to the vitagene family can be extended. In particular, it seems likely that SOD, a major inducible enzyme of the first level of antioxidant defence, can meet selecting criteria to be included in the vitagene family. The products of the aforementioned genes actively operate in detecting and controlling diverse forms of stress and cell injuries. The cooperative mechanisms of the vitagene network are reviewed in recently published comprehensive reviews [234,237,238] with a major conclusion indicating an essential regulatory role of the vitagene network in cell and whole organism adaptation to various stresses.
Emerging findings suggest a large number of potential mechanisms of action of SM (polyphenols) in preventing disease, which may be beyond their direct conventional antioxidant activities. It seems likely that SM, similar to other flavonoids, can affect the vitagene network. The recent relevant findings are reviewed below. First of all, SM/silybin affects HO-1 activity in different model systems. For example, As-intoxicated rats showed a significant up-regulation of myocardial NADPH (NOX) oxidase sub-units such as NOX2 and NOX4 as well as Keap1 and down-regulation of Nrf2 and vitagene HO-1 protein expressions. Pre-administration of silibinin (75 mg/kg/BW) recovered all these altered parameters to near normalcy in As-induced cardiotoxic rat [71]. Similarly, in a model of liver injury caused by alcohol plus pyrazole, SM administration (50 mg/kg/BW) had a protective effect with a trend in restoring the decreased activity of HO-1 and Nrf2 [183]. SM (250 mg/kg/BW) possesses substantial protective effect against B(a)P-induced damages by increasing (restoring) HO-1 (vitagene) activity [115]. Similarly, in vitro SM (500 μM) reduced tBH-induced hepatocyte toxicity by activating HO-1 gene expression [239]. Indeed, the enzyme HO-1 is an important regulatory molecule present in most mammalian cells. In fact, the main function of HO-1 is to break down the pro-oxidant molecule heme into three products; carbon monoxide (CO), biliverdin and free iron and actively participate in the antioxidant defence in the human/animal body [240]. Indeed, HO-1 is a stress-inducible protein and can be induced by various oxidative and inflammatory signals. Therefore, HO-1 expression is regarded as an important adaptive cellular mechanism of AO defence in stress conditions [241,242].
Table 1. Major components of the vitagene network (adapted from [226] and [228]).
Table 1. Major components of the vitagene network (adapted from [226] and [228]).
Molecular LevelCellular Level
AO defenseCell proliferation
DNA-repair systemsCell differentiation
Transfer of genetic informationStability of cell membrane
Stress protein synthesisStability of intracellular milieu
Proteosomal function Macromolecular turnover
Tissue and Organ LevelPhysiological and Redox Control Level
Neutralization and removing toxic chemicalsNeuronal response and synaptic plasticity
Tissue regeneration and wound healingStress response
Tumor suppressionHormonal response
Cell death and cell replacementImmune response
Thermoregulation
HO-1/CO; BVR/BR; UCP; Hsp70; Hsp27; TRXr/TRX
From the data presented above it is clear that SM/silibinin can upregulate HO-1 and improve antioxidant defences. Secondly, SM/silibinin can affect other Hsps including Hsp70. Indeed, in an in vitro system based on CHO-K1 cells treated with As, SM (5 μM) significantly decreased Hsp70 expression previously elevated by As [243]. In another in vitro system based on heat-induced chicken hepatocytes, SM (259 μM) affected Hsp70 expression significantly, preventing its alleviation by heat stress [218]. A similar protective effect of SM (100 mg/kg/BW) on Hsp70 was seen in rats given SM for 7 days prior to mesenteric ischemia-reperfusion (I-R) compared to I-R group [244]. It is interesting to note that silybin was identified as a novel Hsp90 inhibitor [245]. Therefore, silibinin can decrease Hsp70 expression in stressed cells indicating improved AO defences and decrease stress by other means (e.g., Nrf2-related increased AO synthesis).
Thirdly, there are several recent publications showing a regulatory role of SM/silibinin on the sirtuin functions. In fact, a protective effect of silibinin against chemotherapeutic reagent mitomycin C-induced cell death in A375-S2 cells was studied. It was shown that mitomycin C caused cell apoptosis and over-expression of p53, elevated translocation of p53 into the nucleus, and decreased SIRT1 expression; and these effects were alleviated by silibinin (100–150 μM) to a large extent [246]. Silibinin protected cardiac myocytes against isoproterenol-induced injury through resuming mitochondrial function and regulating the expression of SIRT1 and Bcl-2 family members [247]. Systemic administration of silibinin (50 mg/kg/BW) reversed streptozotocin-induced downregulation of SIRT1 expression in mice [248]. It is important to note that SIRT1 was activated by SM (500 μM) in UV-irradiated human malignant melanoma, A375-S2 cells [249]. However, silibinin could have completely the opposite action in regulating SIRT1 expression in normal cardiac myocytes and breast cancer MCF-7 cells, where it downragulates SIRT1 and promotes apoptosis [250]. Similarly silybin treatment downregulated SIRT1 in human lung adenocarcinoma cells [251].
Recent findings indicate that sirtuins-mediated signaling pathways are involved in metabolic reprogramming and upregulation of antioxidant defense systems [252]. Sirtuins are a class of proteins possessing deacetylase activity and playing critical roles in cell survival in response to oxidative stress. Seven human sirtuins, SIRT1–SIRT7, have been identified and their localization differ considerably: SIRT1 and SIRT2- nucleus and cytoplasm; SIRT3, SIRT4 and SIRT5- mitochondria; SIRT3- nucleus and cytoplasm and SIRT6 and SIRT7- nucleus [253]. Sirtuins control DNA repair and recombination, chromosomal stability and gene transcription, as well as regulate antioxidant defences. On the one hand, sirtuins decrease ROS levels, inhibit apoptosis and decrease inflammation. On the other, oxidative stress has been shown to decrease SIRT1 expression in the rat hippocampus and cortex, while SIRT1 overexpression prevents oxidative stress-induced apoptosis and increases resistance to oxidative stress (for review see [254]). Sirtuins are responsible for deacetylation of important transcription factors such as p53, FoxO, NF-κB, or PAPARGCα, which control the transcription of pro- and antioxidant enzymes [255]. Therefore, SM-related activation or prevention of inhibition of sirtuins in stress conditions could be considered as an important adaptive mechanism responsible for maintaining redox-regulated homeostasis in the cell and in the whole body. Finally, there is only a single publication devoted to the protective effect of SM on another vitagene called thioredoxins system. Silybin-phospholipid complex containing vitamin E (Realsil®, Indena, IBI-Lorenzini spa, Italy) was administered daily by gavage (15 mg vitamin E and 47 mg silybin complexed with phospholipids) to rats fed a choline-deprived or a high fat diet for 30 days and 60 days, respectively. Realsil was shown to be able to maintain Trx concentration in the rat liver fed on a high fat diet at day 60, while in the control group it dramatically decreased [59]. It should be mentioned that Trx, a small, ubiquitous thiol protein, is one of the most important regulators of redox balance and redox-sensitive cell functions. In fact, the thioredoxin system consisting of Trx, thioredoxin reductase (TrxR), and NADPH, is a key element determining cell redox status and regulating protein dithiol/disulfide balance, signal transduction and gene expression [256]. Indeed, the Trx system together with the glutathione-glutaredoxin (Grx) system (NADPH, glutathione reductase, GSH, and Grx) controls the cellular redox environment and determines a variety of cellular functions including DNA metabolism and repair, transcription, intracellular signaling, cell-cell communication, cell growth, apoptosis inhibition and ultimately cell survival in various stress conditions [256,257]. It is well accepted that the ability of SM to maintain Trx system in stress conditions is an important element of improving the adaptive ability of animals/humans via the vitagene network. Therefore, the vitagene network represents the major cellular pathway involved in the so called “programmed cell life” as an opposite to apoptosis, providing an effective protection against oxidative stress and toxic products of ROS metabolism [230,231,232]. In particular, nutritional antioxidants can be neuroprotective through the activation of hormetic pathways under the control of the Vitagene protein network [234]. The aforementioned results clearly indicate that, similar to other flavonoids SM, or its active ingredient silibinin, affect vitagene network decreasing detrimental consequences of stresses. Further research is needed to fully understand the molecular mechanisms involved in the vitagene network activation by SM/silybin.

7. Protective Effect of SM in the Gut

It seems likely that the gut is the major place of antioxidant action of polyphenols [258,259]. Indeed, reduction of oxidative damage, modulation of colonic flora and variation in gene expression are involved in the modulation of intestinal function by polyphenols. For example, to study the molecular effects of wine polyphenols at the gene level, the microarray technology was used: rats were treated with 50 mg/kg/BW wine polyphenols for 14 days. Global expression analysis of 5707 genes revealed an extensive down-regulation of genes involved in a wide range of physiological functions, such as metabolism, transport, signal transduction and intercellular signaling [260]. It was shown that two major regulatory pathways were down-regulated in the colon mucosa of polyphenols treated rats: inflammatory response and steroid metabolism. Since flavonoids are consumed in concentrations usually much higher than other antioxidant compounds, their protective effect during digestion is of great importance. For example, it has been shown that flavonoids not only prevented an accumulation of peroxidized lipids but can also switch prooxidant properties of heme-proteins to antioxidant ones [261]. Dietary polyphenols can also modulate in vivo oxidative damage in the gastrointestinal tract of rodents [262] supporting the hypothesis that dietary polyphenols might have both a protective and a therapeutic potential in oxidative damage-related pathologies. Indeed, the antioxidant-prooxidant balance (redox status) in various parts of the intestine would ultimately depend on the level of antioxidants and pro-oxidants provided by the diet and released by cells themselves, as well as on the level of absorption of both antioxidants and pro-oxidants [258,259]. In a model system mimicking stomach conditions it was shown that both lipid peroxidation and co-oxidation of vitamin E and beta-carotene were inhibited at pH 3.0 by red wine polyphenols [263]. In the gut, an interaction between different antioxidants could bring about their maximum synergistic effects. In particular, it was found that the mixture of the green tea polyphenol, vitamin E and vitamin C could act synergistically to protect lipid peroxidation [264]. It has been found that consumption of partially oxidized food could increase lipid peroxidation in the stomach and the absorption of cytotoxic lipid peroxidation products into the body. The addition of red wine polyphenols to the meal may alter these outcomes [265], therefore red wine polyphenols exert a beneficial effect by the novel function, absorption inhibition of the MDA [266]. These findings explain the potentially harmful effects of oxidized fats found in foods/feeds [259] and the important benefit of dietary polyphenols in the diet. Indeed, locating the biological site of action of polyphenols in the gastrointestinal tract may lead to a revision in our understanding of how flavonoids, including silybin, work in vivo and may help to elucidate the mechanisms of the health-promoting action of silybin and other polyphenols [267].
Redox signaling in gut inflammation is complex and poorly understood. However, it is generally accepted that homeostatic control of the intestinal epithelial redox environment is central for nutrient digestion and absorption, stem cell proliferation, apical enterocyte apoptosis, and immune response [268]. Indeed, polyphenols may play a role in intestinal mucosa integrity, inflammation and permeability [258,269]. For example, wine phenolics were able to prevent or delay the progression of intestinal diseases characterized by oxidative stress and inflammation, acting as both free radical scavengers and modulators of specific inflammation-related genes involved in cellular redox signaling [270]. They exert their effects by modulating cell signaling pathways, mainly activated in response to oxidative and inflammatory stimuli, and Nrf2 and NF-κB are the principal downstream effectors [271]. For example, pigs administered a grape seed and grape marc extract had a lower transactivation of NF-κB and Nrf2 and a lower expression of various target genes of these transcription factors in the duodenal mucosa (but not in the liver and plasma) than control pigs [272]. Furthermore, the ratio of villus height: crypt depth and the gain:feed ratio was higher in the animals fed a gape extract than in control pigs. In mice, the dietary grape seed extract decreased proliferation and enhanced differentiation of epithelial cell and the changes in gut epithelium were associated with the suppression of NF-κB signaling [273]. Polyphenolic compounds could also affect gut integrity. For example, grape seed extract (0.1%) supplied to rats with drinking water for 21 days significantly increased the expression of gut junction protein occludin in the proximal colon and reduced fecal levels of the neutrophil protein calprotectin, compared with control [274]. Recently, it has been reported that a naturally occurring flavonoid procyanidin B2 could modulate cellular redox status and the antioxidant enzyme defence system in colonic cells protecting against oxidative stress and xenobiotics [275,276].
It is possible to suggest that there is a biological reason for some antioxidants not to be absorbed completely and so provide antioxidant protection in lower parts of the intestine [258,259]. Comparatively low bioavailability and antioxidant potential of various flavonoids could be beneficial for humans/animals providing antioxidant protection in various parts of the digestive tract, including the large intestine where levels of other antioxidants would be quite low. The same is true for SM. For example, silibinin concentration in the gut could reach 800 μM [277], a concentration exceeding the requirement for its direct antioxidant activity. Indeed, it is still unclear whether SM has any direct antioxidant effects in vivo, although it might be capable of exerting such effects within the gastrointestinal tract, where SM may come into direct contact with cells without having undergone absorption and metabolism [258]. It has been suggested that gastroprotective effects of SM might be related to the protection of gastric mucosal NO and non-protein sulfhydryl groups and the modulation of capsaicin-sensitive gastric sensory afferents [278]. The beneficial effects of SM, on a murine model of colitis was explained [279]. SM effectively reduced colonic damage, lipid peroxidation, neutrophil infiltration and the content of inflammatory cytokines alongside increasing total antioxidant capacity of colonic tissue. In trinitrobenzene sulfonic acid-induced colitis model, silibinin treatment significantly reduced several components of inflammatory colitis such as NF-κB activity, levels of IL-1β, TNFα, TBARS, protein carbonyl, myeloperoxidase activity, and an improvement in antioxidant capability of the colon tissue [280]. Furthermore, silibinin (50 mg/kg/BW) modulates gut microbial enzymes, colonic oxidative stress and Wnt/β-catenin signaling resulting in antiproliferative effect against 1,2-di-methylhydrazine-induced colon carcinogenesis in experimental rats [64].
It is also well known that polyphenols are providing antioxidant protection in the lower intestine and can alter colonic flora [281]. Polyphenols, including silybin, are extensively metabolized by gut bacteria into a complex series of end-products that affect the functional ecology of symbiotic partners that can alter the host physiology [281]. It has been suggested that the consumption of a diet rich in plant foods with high dietary polyphenol content may enhance the gastrointestinal health of the host through microbiota modulation. Indeed, grape antioxidant dietary fiber intake stimulates proliferation of Lactobacillus and slightly affects the composition of Bifidobacterium species in the cecum of rats [282]. It appears that dietary polyphenols may have the ability to modify the gut microbial balance, but this effect is indirect, i.e., it is mediated by biotransformation products, rather than the original plant compounds [283]. It should be mentioned that polyphenol–microbiota interactions are complex and subject to large inter-individual variability, leading to different polyphenol-metabolizing phenotypes or “metabotypes” [284].
Unfortunately, there is no data available on the effect of SM on gut microbiota, but we can suggest the same changes would occur as described above for other flavonoids. Therefore, it seems likely that reduction of oxidative damage, decreased inflammation, modulation of colonic flora and variation in gene expression are involved in the modulation of intestinal function by polyphenols, including SM.

8. Conclusions

Recent achievements in biochemistry and molecular biology, together with epidemiological data have changed our thinking about food. It has become increasingly clear that our diet plays a pivotal role in maintenance of our health and a misbalanced diet can cause serious health-related problems. It seems likely that antioxidants are among the major regulators of many physiological processes and, therefore, a redox balance between antioxidants and prooxidants in the diet, gastro-intestinal tract, plasma and tissues is an important determinant of the state of our health. Plants consumed by humans amd animals contain thousands of phenolic compounds. Among them, the effects of dietary polyphenols including SM are of great current interest. Indeed, various phytochemicals, including flavonoids, are an essential part of our diet, which are responsible for turning on and maintaining an optimal status of our antioxidant defences. Since flavonoids are not well absorbed in the gut, their active concentration in the plasma and target tissues are comparatively low, but probably sufficient for Nrf2 activation and NF-κB suppression as well as vitagene activation. Indeed, it seems very likely that activation of the Keap1/Nrf2/ARE pathway and inhibition of NF-κB pathway, rather than direct free radical scavenging activity, may be the main mechanisms of the health benefits of phytochemicals [233], including SM. Therefore, consumption of phytochemicals, including SM, could have a pre-conditioning effect on the antioxidant system of the body. This could explain the beneficial health-promoting effects of a diet rich in fruits and vegetables as important sources of the aforementioned chemicals (polyphenols and other phytochemicals) maintaining the body’s ability to be highly adaptive to various stresses. SM and its main component silibinin are part of the dietary phytochemical mixture responsible for regulation of the antioxidant defences in the gut and in the whole body. It could well be that some dietary constituents which are not well absorbed could have health-promoting properties by maintaining redox balance in the large intestine, where concentration of other antioxidants (vitamin E, carotenoids, ascorbate) could be low, but prooxidants (iron, oxidized PUFAs, etc.) and substrates of oxidation are still present [258,259]. This protective effect in the large intestine could be responsible, for example, for bowl cancer prevention. Therefore, there could be a biological reason for some nutrients not being absorbed, but still being involved in antioxidant protection in the lower gut. Taking into account high concentrations of phytochemicals in the gut, it could well be that they play an essential part in maintaining an optimal antioxidant-prooxidant balance in the digestive tract responsible for additional health effects of phytochemicals including SM. In animal nutrition and disease prevention strategy SM alone, or in combination with other hepatho-active compomds (carnitine, betaine, vitamin B12, etc.), could have similar hepatoprotective effects as described in humans with similar mechanisms of protective action.
In conclusion, there are many possible mechanisms by which SM can improve the antioxidant defence mechanisms in the body. They include direct and indirect SM actions. First of all, a direct scavenging free radicals and chelating free Fe and Cu are mainly effective in the gut. Secondly, preventing free radical formation by inhibiting specific ROS-producing enzymes, or improving the integrity of electron-transport chain of mitochondria in stress conditions as a result of SM consumption, is of great importance. Thirdly, maintaining an optimal redox balance in the cell by activating a range of antioxidant enzymes and non-enzymatic antioxidants, mainly via Nrf2 activation, is probably the main driving force of AO action of SM. Fourthly, decreasing inflammatory responses in the gut and other tissues by inhibiting NF-kB pathways is an emerging mechanism of SM protective effect in liver toxicity and diseases. Fifthly, activating vitagenes, responsible for synthesis of protective molecules, including HSP, Trx, sirtuins, etc., and providing additional protection in stress conditions deserves more attention in future research. Finally, affects on the microenvironment of the gut, including SM-bacteria interactions, await future investigation.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Comelli, M.C.; Mengs, U.; Prosdocimi, M.; Schneider, C. Toward the definition of the mechanism of action of silymarin: Activities related to cellular protection from toxic damage induced by chemotherapy. Integr. Cancer Ther. 2007, 6, 120–129. [Google Scholar] [CrossRef] [PubMed]
  2. Biedermann, D.; Vavříková, E.; Cvak, L.; Křen, V. Chemistry of silybin. Nat. Prod. Rep. 2014, 31, 1138–1157. [Google Scholar] [CrossRef] [PubMed]
  3. Křen, V.; Marhol, P.; Purchartová, K.; Gabrielová, E.; Modrianský, M. Biotransformation of silybin and its congeners. Curr. Drug MeTable 2013, 14, 1009–1021. [Google Scholar] [CrossRef]
  4. Hackett, E.S.; Twedt, D.C.; Gustafson, D.L. Milk thistle and its derivative compounds: A review of opportunities for treatment of liver disease. J. Vet. Intern. Med. 2013, 27, 10–16. [Google Scholar] [CrossRef] [PubMed]
  5. Post-White, J.; Ladas, E.J.; Kelly, K.M. Advances in the use of milk thistle (Silybum marianum). Integr. Cancer Ther. 2007, 6, 104–109. [Google Scholar] [CrossRef] [PubMed]
  6. Loguercio, C.; Federico, A.; Trappoliere, M.; Tuccillo, C.; de Sio, I.; di Leva, A.; Niosi, M.; D’Auria, M.V.; Capasso, R.; del Vecchio Blanco, C.; et al. The effect of a silybin-vitamin E-phospholipid complex on nonalcoholic fatty liver disease: A pilot study. Dig. Dis. Sci. 2007, 52, 2387–2395. [Google Scholar] [CrossRef] [PubMed]
  7. Federico, A.; Niosi, M.; del Vecchio Blanco, C.; Loguercio, C. Emerging drugs for non-alcoholic fatty liver disease. Exp. Opin. Emerg. Drugs 2008, 13, 145–158. [Google Scholar] [CrossRef]
  8. Trappoliere, M.; Caligiuri, A.; Schmid, M.; Bertolani, C.; Failli, P.; Vizzutti, F.; Novo, E.; di Manzano, C.; Marra, F.; Loguercio, C.; et al. Silybin, a component of sylimarin, exerts anti-inflammatory and anti-fibrogenic effects on human hepatic stellate cells. J. Hepatol. 2009, 50, 1102–1111. [Google Scholar] [CrossRef] [PubMed]
  9. Loguercio, C.; Festi, D. Silybin and the liver: From basic research to clinical practice. World J. Gastroenterol. 2011, 17, 2288–2301. [Google Scholar] [CrossRef] [PubMed]
  10. Loguercio, C.; Andreone, P.; Brisc, C.; Brisc, M.C.; Bugianesi, E.; Chiaramonte, M.; Cursaro, C.; Danila, M.; de Sio, I.; Floreani, A.; et al. Silybin combined with phosphatidylcholine and vitamin E in patients with nonalcoholic fatty liver disease: A randomized controlled trial. Free Radic. Biol. Med. 2012, 52, 1658–1665. [Google Scholar] [CrossRef] [PubMed]
  11. Stiuso, P.; Scognamiglio, I.; Murolo, M.; Ferranti, P.; de Simone, C.; Rizzo, M.R.; Tuccillo, C.; Caraglia, M.; Loguercio, C.; Federico, A. Serum oxidative stress markers and lipidomic profile to detect NASH patients responsive to an antioxidant treatment: A pilot study. Oxid. Med. Cell Longev. 2014, 2014, 169216. [Google Scholar] [CrossRef] [PubMed]
  12. Testino, G.; Leone, S.; Ansaldi, F.; Borro, P. Silymarin and S-adenosyl-l-methionine (SAMe): Two promising pharmacological agents in case of chronic alcoholic hepathopathy. A review and a point of view. Minerva Gastroenterol. Dietol. 2013, 59, 341–356. [Google Scholar] [PubMed]
  13. Milić, N.; Milosević, N.; Suvajdzić, L.; Zarkov, M.; Abenavoli, L. New therapeutic potentials of milk thistle (Silybum marianum). Nat. Prod. Commun. 2013, 8, 1801–1810. [Google Scholar] [PubMed]
  14. Zholobenko, A.; Modriansky, M. Silymarin and its constituents in cardiac preconditioning. Fitoterapia 2014, 97, 122–132. [Google Scholar] [CrossRef] [PubMed]
  15. Madrigal-Santillán, E.; Madrigal-Bujaidar, E.; Álvarez-González, I.; Sumaya-Martínez, M.T.; Gutiérrez-Salinas, J.; Bautista, M.; Morales-González, Á.; García-Luna, Y.; González-Rubio, M.; Aguilar-Faisal, J.L.; et al. Review of natural products with hepatoprotective effects. World J. Gastroenterol. 2014, 20, 14787–14804. [Google Scholar] [CrossRef] [PubMed]
  16. Vargas-Mendoza, N.; Madrigal-Santillán, E.; Morales-González, A.; Esquivel-Soto, J.; Esquivel-Chirino, C.; García-Luna, Y.; González-Rubio, M.; Gayosso-de-Lucio, J.A.; Morales-González, J.A. Hepatoprotective effect of silymarin. World J. Hepatol. 2014, 6, 144–149. [Google Scholar] [CrossRef] [PubMed]
  17. Mohd Fozi, N.F.; Mazlan, M.; Shuid, A.N.; Isa Naina, M. Milk thistle: A future potential anti-osteoporotic and fracture healing agent. Curr. Drug Targets 2013, 14, 1659–1666. [Google Scholar] [CrossRef] [PubMed]
  18. Surai, P.F. Polyphenol compounds in the chicken/animal diet: From the past to the future. J. Anim. Physiol. Anim. Nutr. (Berl.) 2014, 98, 19–31. [Google Scholar] [CrossRef]
  19. Hollman, P.C. Unravelling of the health effects of polyphenols is a complex puzzle complicated by metabolism. Arch. Biochem. Biophys. 2014, 559, 100–105. [Google Scholar] [CrossRef] [PubMed]
  20. Li, L.; Zeng, J.; Gao, Y.; He, D. Targeting silibinin in the antiproliferative pathway. Expert Opin. Investig. Drugs 2010, 19, 243–255. [Google Scholar] [CrossRef] [PubMed]
  21. Wen, Z.; Dumas, T.E.; Schrieber, S.J.; Hawke, R.L.; Fried, M.W.; Smith, P.C. Pharmacokinetics and metabolic profile of free, conjugated, and total silymarin flavonolignans in human plasma after oral administration of milk thistle extract. Drug Metab. Dispos. 2008, 36, 65–72. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, H.J.; Brinda, B.J.; Chavin, K.D.; Bernstein, H.J.; Patrick, K.S.; Markowitz, J.S. An assessment of pharmacokinetics and antioxidant activity of free silymarin flavonolignans in healthy volunteers: A dose escalation study. Drug Metab. Dispos. 2013, 41, 1679–1685. [Google Scholar] [CrossRef] [PubMed]
  23. Wu, J.W.; Lin, L.C.; Hung, S.C.; Chi, C.W.; Tsai, T.H. Analysis of silibinin in rat plasma and bile for hepatobiliary excretion and oral bioavailability application. J. Pharm. Biomed. Anal. 2007, 45, 635–641. [Google Scholar] [CrossRef] [PubMed]
  24. Lorenz, D.; Lucker, P.W.; Mennicke, W.H.; Wetzelsberger, N. Pharmacokinetic studies with silymarin in human serum and bile. Methods Find. Exp. Clin. Pharmacol. 1984, 6, 655–661. [Google Scholar] [PubMed]
  25. Barzaghi, N.; Crema, F.; Gatti, G.; Pifferi, G.; Perucca, E. Pharmokinetic studies in IdB 1016, a silybin-phosphatidylcholine complex, in healthy human subjects. Eur. J. Drug. Metab. Pharmacokinet. 1990, 15, 333–338. [Google Scholar] [CrossRef] [PubMed]
  26. Flory, P.J.; Krug, G.; Lorenz, D.; Mennicke, W.H. Studies on elimination of silymarin in cholecystectomized patients. I. Biliary and renal elimination after a single oral dose. Plant. Med. 1980, 38, 227–237. [Google Scholar] [CrossRef]
  27. Zhao, J.; Agarwal, R. Tissue distribution of silibinin, the major active constituent of silymarin, in mice and its association with enhancement of phase II enzymes: Implications in cancer chemoprevention. Carcinogenesis 1999, 20, 2101–2108. [Google Scholar] [CrossRef] [PubMed]
  28. Han, Y.H.; Lou, H.X.; Ren, D.M.; Sun, L.R.; Ma, B.; Ji, M. Stereoselective metabolism of silybin diastereoisomers in the glucuronidation process. J. Pharm. Biomed. Anal. 2004, 34, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  29. D’Andrea, V.; Perez, L.M.; Sanchez Pozzi, E.J. Inhibition of rat liver UDP glucuronosyltransferase by silymarin and the metabolite silibinin-glucuronide. Life Sci. 2005, 77, 683–692. [Google Scholar] [CrossRef] [PubMed]
  30. Weyhenmeyer, R.; Mascher, H.; Birkmayer, J. Study on dose-linearity of the pharmacokinetics of silibinin diastereomers using a new stereospecific assay. Int. J. Clin. Pharmacol. Ther. Toxicol. 1992, 30, 134–138. [Google Scholar] [PubMed]
  31. Calani, L.; Brighenti, F.; Bruni, R.; del Rio, D. Absorption and metabolism of milk thistle flavanolignans in humans. Phytomedicine 2012, 20, 40–46. [Google Scholar] [CrossRef] [PubMed]
  32. Rickling, B.; Hans, B.; Kramarczyk, R.; Krumbiegel, G.; Weyhenmeyer, R. Two high-performance liquid chromatographic assays for the determination of free and total silibinin diastereomers in plasma using column switching with electrochemical detection and reversed-phase chromatography with ultraviolet detection. J. Chromatogr. B Biomed. Appl. 1995, 670, 267–277. [Google Scholar] [CrossRef] [PubMed]
  33. Surai, P.F.; Fisinin, V.I. Antioxidant Systems of the Body: From Vitamin E to Polyphenols and Beyond. In Proceedings of the 35th Western Nutrition Conference, Edmonton, Canada, 24–25 September 2014; pp. 265–277.
  34. Surai, P.F. Natural Antioxidants in Avian Nutrition and Reproduction; Nottingham University Press: Nottingham, UK, 2002. [Google Scholar]
  35. Surai, P.F. Selenium in Nutrition and Health; Nottingham University Press: Nottingham, UK, 2006. [Google Scholar]
  36. Surai, P.F.; Fisinin, V.I. Ill Health effects of food lipids: Consequences of inadequate food processing, storage and cooking. In Modern Dietary Fat Intakes in Disease Promotion; de Meester, F., Zibadi, S., Watson, R.R., Eds.; Springer: New York, NY, USA, 2010; pp. 251–274. [Google Scholar]
  37. Fotina, A.A.; Fisinin, V.I.; Surai, P.F. Recent developments in usage of natural antioxidants to improve chicken meat production and quality. Bulg. J. Agric. Sci. 2013, 19, 889–896. [Google Scholar]
  38. Dehmlow, C.; Erhard, J.; de Groot, H. Inhibition of Kupffer cell functions as an explanation for the hepatoprotective properties of silibinin. Hepatology 1996, 23, 749–754. [Google Scholar] [CrossRef] [PubMed]
  39. Dehmlow, C.; Murawski, N.; de Groot, H. Scavenging of reactive oxygen species and inhibition of arachidonic acid metabolism by silibinin in human cells. Life Sci. 1996, 58, 1591–1600. [Google Scholar] [CrossRef] [PubMed]
  40. Varga, Z.; Seres, I.; Nagy, E.; Ujhelyi, L.; Balla, G.; Balla, J.; Antus, S. Structure prerequisite for antioxidant activity of silybin in different biochemical systems in vitro. Phytomedicine 2006, 13, 85–93. [Google Scholar] [CrossRef] [PubMed]
  41. Fu, H.; Lin, M.; Muroya, Y.; Hata, K.; Katsumura, Y.; Yokoya, A.; Shikazono, N.; Hatano, Y. Free radical scavenging reactions and antioxidant activities of silybin: Mechanistic aspects and pulse radiolytic studies. Free Radic. Res. 2009, 43, 887–897. [Google Scholar] [CrossRef] [PubMed]
  42. Yin, F.; Liu, J.; Ji, X.; Wang, Y.; Zidichouski, J.; Zhang, J. Silibinin: A novel inhibitor of Aβ aggregation. Neurochem. Int. 2011, 58, 399–403. [Google Scholar] [CrossRef] [PubMed]
  43. Cristofalo, R.; Bannwart-Castro, C.F.; Magalhães, C.G.; Borges, V.T.; Peraçoli, J.C.; Witkin, S.S.; Peraçoli, M.T. Silibinin attenuates oxidative metabolism and cytokine production by monocytes from preeclamptic women. Free Radic. Res. 2013, 47, 268–275. [Google Scholar] [CrossRef] [PubMed]
  44. Domitrović, R.; Jakovac, H.; Marchesi, V.V.; Blažeković, B. Resolution of liver fibrosis by isoquinoline alkaloid berberine in CCl4-intoxicated mice is mediated by suppression of oxidative stress and upregulation of MMP-2 expression. J. Med. Food 2013, 16, 518–528. [Google Scholar] [CrossRef] [PubMed]
  45. Köksal, E.; Gülçin, I.; Beyza, S.; Sarikaya, O.; Bursal, E. In vitro antioxidant activity of silymarin. J. Enzyme Inhib. Med. Chem. 2009, 24, 395–405. [Google Scholar] [CrossRef]
  46. Asghar, Z.; Masood, Z. Evaluation of antioxidant properties of silymarin and its potential to inhibit peroxyl radicals in vitro. Pak. J. Pharm. Sci. 2008, 21, 249–254. [Google Scholar] [PubMed]
  47. Dvorak, Z.; Kosina, P.; Walterova, D.; Simanek, V.; Bachleda, P.; Ulrichova, J. Primary cultures of human hepatocytes as a tool in cytotoxicity studies: Cell protection against model toxins by flavonolignans obtained from Silybum marianum. Toxicol. Lett. 2003, 137, 201–212. [Google Scholar] [CrossRef] [PubMed]
  48. Crocenzi, F.A.; Mottino, A.D.; Roma, M.G. Regulation of synthesis and trafficking of canalicular transporters and its alteration in acquired hepatocellular cholestasis. Experimental therapeutic strategies for its prevention. Curr. Med. Chem. 2004, 11, 501–524. [Google Scholar] [CrossRef] [PubMed]
  49. Blaising, J.; Lévy, P.L.; Gondeau, C.; Phelip, C.; Varbanov, M.; Teissier, E.; Ruggiero, F.; Polyak, S.J.; Oberlies, N.H.; Ivanovic, T.; et al. Silibinin inhibits hepatitis C virus entry into hepatocytes by hindering clathrin-dependent trafficking. Cell Microbiol. 2013, 15, 1866–1882. [Google Scholar] [PubMed]
  50. Colturato, C.P.; Constantin, R.P.; Maeda, A.S., Jr.; Constantin, R.P.; Yamamoto, N.S.; Bracht, A.; Ishii-Iwamoto, E.L.; Constantin, J. Metabolic effects of silibinin in the rat liver. Chem. Biol. Interact. 2012, 195, 119–132. [Google Scholar] [CrossRef] [PubMed]
  51. Lin, M.T.; Beal, M.F. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature 2006, 443, 787–795. [Google Scholar] [CrossRef] [PubMed]
  52. Sekine, S.; Ichijo, H. Mitochondrial proteolysis: Its emerging roles in stress responses. Biochim. Biophys. Acta 2015, 1850, 274–280. [Google Scholar] [CrossRef] [PubMed]
  53. Calabrese, V.; Cornelius, C.; Stella, A.M.; Calabrese, E.J. Cellular stress responses, mitostress and carnitine insufficiencies as critical determinants in aging and neurodegenerative disorders: Role of hormesis and vitagenes. Neurochem. Res. 2010, 35, 1880–1915. [Google Scholar] [CrossRef] [PubMed]
  54. Rolo, A.P.; Oliveira, P.J.; Moreno, A.J.; Palmeira, C.M. Protection against post-ischemic mitochondrial injury in rat liver by silymarin or TUDC. Hepatol. Res. 2003, 26, 217–224. [Google Scholar] [CrossRef] [PubMed]
  55. Zhou, B.; Wu, L.J.; Tashiro, S.; Onodera, S.; Uchiumi, F.; Ikejima, T. Silibinin protects rat cardiac myocyte from isoproterenol-induced DNA damage independent on regulation of cell cycle. Biol. Pharm. Bull. 2006, 29, 1900–1905. [Google Scholar] [CrossRef] [PubMed]
  56. Detaille, D.; Sanchez, C.; Sanz, N.; Lopez-Novoa, J.M.; Leverve, X.; El-Mir, M.Y. Interrelation between the inhibition of glycolytic flux by silibinin and the lowering of mitochondrial ROS production in perifused rat hepatocytes. Life Sci. 2008, 82, 1070–1076. [Google Scholar] [CrossRef] [PubMed]
  57. Ligeret, H.; Brault, A.; Vallerand, D.; Haddad, Y.; Haddad, P.S. Antioxidant and mitochondrial protective effects of silibinin in cold preservation-warm reperfusion liver injury. J. Ethnopharmacol. 2008, 115, 507–514. [Google Scholar] [CrossRef] [PubMed]
  58. Gabrielová, E.; Jabůrek, M.; Gažák, R.; Vostálová, J.; Ježek, J.; Křen, V.; Modrianský, M. Dehydrosilybin attenuates the production of ROS in rat cardiomyocyte mitochondria with an uncoupler-like mechanism. J. Bioenerg. Biomembr. 2010, 42, 499–509. [Google Scholar] [CrossRef] [PubMed]
  59. Grattagliano, I.; Diogo, C.V.; Mastrodonato, M.; de Bari, O.; Persichella, M.; Wang, D.Q.; Liquori, A.; Ferri, D.; Carratù, M.R.; Oliveira, P.J.; et al. A silybin-phospholipids complex counteracts rat fatty liver degeneration and mitochondrial oxidative changes. World J. Gastroenterol. 2013, 19, 3007–3017. [Google Scholar] [CrossRef] [PubMed]
  60. Serviddio, G.; Bellanti, F.; Giudetti, A.M.; Gnoni, G.V.; Petrella, A.; Tamborra, R.; Romano, A.D.; Rollo, T.; Vendemiale, G.; Altomare, E. A silybin-phospholipid complex prevents mitochondrial dysfunction in a rodent model of nonalcoholic steatohepatitis. J. Pharmacol. Exp. Ther. 2010, 332, 922–932. [Google Scholar] [CrossRef] [PubMed]
  61. Zhu, S.Y.; Dong, Y.; Tu, J.; Zhou, Y.; Zhou, X.H.; Xu, B. Silybum marianum oil attenuates oxidative stress and ameliorates mitochondrial dysfunction in mice treated with d-galactose. Pharmacogn. Mag. 2014, 10, S92–S99. [Google Scholar] [CrossRef] [PubMed]
  62. Geed, M.; Garabadu, D.; Ahmad, A.; Krishnamurthy, S. Silibinin pretreatment attenuates biochemical and behavioral changes induced by intrastriatal MPP+ injection in rats. Pharmacol. Biochem. Behav. 2014, 117, 92–103. [Google Scholar] [CrossRef] [PubMed]
  63. Mazzio, E.A.; Harris, N.; Soliman, K.F. Food constituents attenuate monoamine oxidase activity and peroxide levels in C6 astrocyte cells. Plant. Med. 1998, 64, 603–606. [Google Scholar] [CrossRef]
  64. Sangeetha, N.; Viswanathan, P.; Balasubramanian, T.; Nalini, N. Colon cancer chemopreventive efficacy of silibinin through perturbation of xenobiotic metabolizing enzymes in experimental rats. Eur. J. Pharmacol. 2012, 674, 430–438. [Google Scholar] [CrossRef] [PubMed]
  65. Tuorkey, M.J.; El-Desouki, N.I.; Kamel, R.A. Cytoprotective Effect of Silymarin against Diabetes-Induced Cardiomyocyte Apoptosis in Diabetic Rats. Biomed. Environ. Sci. 2015, 28, 36–43. [Google Scholar] [PubMed]
  66. Fan, L.; Ma, Y.; Liu, Y.; Zheng, D.; Huang, G. Silymarin induces cell cycle arrest and apoptosis in ovarian cancer cells. Eur. J. Pharmacol. 2014, 743, 79–88. [Google Scholar] [CrossRef] [PubMed]
  67. Doehner, W.; Landmesser, U. Xanthine oxidase and uric acid in cardiovascular disease: Clinical impact and therapeutic options. Semin. Nephrol. 2011, 31, 433–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Varga, Z.; Ujhelyi, L.; Kiss, A.; Balla, J.; Czompa, A.; Antus, S. Effect of silybin on phorbol myristate actetate-induced protein kinase C translocation, NADPH oxidase activity and apoptosis in human neutrophils. Phytomedicine 2004, 11, 206–212. [Google Scholar] [CrossRef] [PubMed]
  69. Pauff, J.M.; Hille, R. Inhibition studies of bovine xanthine oxidase by luteolin, silibinin, quercetin, and curcumin. J. Nat. Prod. 2009, 72, 725–731. [Google Scholar] [CrossRef] [PubMed]
  70. Bedard, K.; Krause, K.H. The NOX family of ROS-generating NADPH oxidases: Physiology and pathophysiology. Physiol. Rev. 2007, 87, 245–313. [Google Scholar] [CrossRef] [PubMed]
  71. Muthumani, M.; Prabu, S.M. Silibinin potentially attenuates arsenic-induced oxidative stress mediated cardiotoxicity and dyslipidemia in rats. Cardiovasc. Toxicol. 2014, 14, 83–97. [Google Scholar] [CrossRef] [PubMed]
  72. Prabu, S.M.; Muthumani, M. Silibinin ameliorates arsenic induced nephrotoxicity by abrogation of oxidative stress, inflammation and apoptosis in rats. Mol. Biol. Rep. 2012, 39, 11201–11216. [Google Scholar] [CrossRef] [PubMed]
  73. Khazim, K.; Gorin, Y.; Cavaglieri, R.C.; Abboud, H.E.; Fanti, P. The antioxidant silybin prevents high glucose-induced oxidative stress and podocyte injury in vitro and in vivo. Am. J. Physiol. Renal. Physiol. 2013, 305, F691–F700. [Google Scholar] [CrossRef] [PubMed]
  74. Ingawale, D.K.; Mandlik, S.K.; Naik, S.R. Models of hepatotoxicity and the underlying cellular, biochemical and immunological mechanism(s): A critical discussion. Environ. Toxicol. Pharmacol. 2014, 37, 118–133. [Google Scholar] [CrossRef] [PubMed]
  75. Rashid, K.; Sinha, K.; Sil, P.C. An update on oxidative stress-mediated organ pathophysiology. Food Chem. Toxicol. 2013, 62, 584–600. [Google Scholar] [CrossRef] [PubMed]
  76. Muthumani, M.; Prabu, S.M. Silibinin potentially protects arsenic-induced oxidative hepatic dysfunction in rats. Toxicol. Mech. Methods 2012, 22, 277–288. [Google Scholar] [CrossRef] [PubMed]
  77. Soria, E.A.; Eynard, A.R.; Bongiovanni, G.A. Modulation of early stress-related biomarkers in cytoplasm by the antioxidants silymarin and quercetin using a cellular model of acute arsenic poisoning. Basic Clin. Pharmacol. Toxicol. 2010, 107, 982–987. [Google Scholar] [CrossRef] [PubMed]
  78. Jain, A.; Yadav, A.; Bozhkov, A.I.; Padalko, V.I.; Flora, S.J. Therapeutic efficacy of silymarin and naringenin in reducing arsenic-induced hepatic damage in young rats. Ecotoxicol. Environ. Saf. 2011, 74, 607–614. [Google Scholar] [CrossRef] [PubMed]
  79. Manibusan, M.K.; Odin, M.; Eastmond, D.A. Postulated carbon tetrachloride mode of action: A review. J. Environ. Sci. Health C Environ. Carcinog. Ecotoxicol. Rev. 2007, 25, 185–209. [Google Scholar] [CrossRef] [PubMed]
  80. Raj, S.; Gothandam, K.M. Hepatoprotective effect of polyphenols rich methanolic extract of Amorphophallus commutatus var. wayanadensis against CCl4 induced hepatic injury in swiss albino mice. Food Chem. Toxicol. 2014, 67, 105–112. [Google Scholar] [PubMed]
  81. Alkreathy, H.M.; Khan, R.A.; Khan, M.R.; Sahreen, S. CCl4 induced genotoxicity and DNA oxidative damages in rats: Hepatoprotective effect of Sonchus arvensis. BMC Complement. Altern. Med. 2014, 14, 452. [Google Scholar] [CrossRef]
  82. Krishnappa, P.; Venkatarangaiah, K.; Shivamogga, V.; Rajanna, S.K.; Prakash, K.; Gupta, R. Antioxidant and prophylactic effects of Delonix elata L., stem bark extracts, and flavonoid isolated quercetin against carbon tetrachloride-induced hepatotoxicity in rats. Biomed. Res. Int. 2014, 2014. [Google Scholar] [CrossRef] [PubMed]
  83. Shaker, E.; Mahmoud, H.; Mnaa, S. Silymarin, the antioxidant component and Silybum marianum extracts prevent liver damage. Food Chem. Toxicol. 2010, 48, 803–806. [Google Scholar] [CrossRef] [PubMed]
  84. Shine, V.J.; Latha, P.G.; Suja, S.N.; Anuja, G.I.; Raj, G.; Rajasekharan, S.N. Ameliorative effect of alkaloid extract of Cyclea peltata (Poir.) Hook. f. & Thoms. Roots (ACP) on APAP/CCl4 induced liver toxicity in Wistar rats and in vitro free radical scavenging property. Asian Pac. J. Trop. Biomed. 2014, 4, 143–151. [Google Scholar] [CrossRef] [PubMed]
  85. Al-Harbi, N.O.; Imam, F.; Nadeem, A.; Al-Harbi, M.M.; Iqbal, M.; Ahmad, S.F. Carbon tetrachloride-induced hepatotoxicity in rat is reversed by treatment with riboflavin. Int. Immunopharmacol. 2014, 21, 383–388. [Google Scholar] [CrossRef] [PubMed]
  86. Chang, H.C.; Chiu, Y.W.; Lin, Y.M.; Chen, R.J.; Lin, J.A.; Tsai, F.J.; Tsai, C.H.; Kuo, Y.C.; Liu, J.Y.; Huang, C.Y. Herbal supplement attenuation of cardiac fibrosis in rats with CCl4-induced liver cirrhosis. Chin. J. Physiol. 2014, 57, 41–47. [Google Scholar] [CrossRef] [PubMed]
  87. Abbas, A.T.; El-Shitany, N.A.; Shaala, L.A.; Ali, S.S.; Azhar, E.I.; Abdel-Dayem, U.A.; Youssef, D.T. Red Sea Suberea mollis Sponge Extract Protects against CCl4-Induced Acute Liver Injury in Rats via an Antioxidant Mechanism. Evid. Based Complement. Alternat. Med. 2014, 2014, 745606. [Google Scholar] [CrossRef] [PubMed]
  88. Li, C.C.; Hsiang, C.Y.; Wu, S.L.; Ho, T.Y. Identification of novel mechanisms of silymarin on the carbon, tetrachloride-induced liver fibrosis in mice by nuclear factor-κB bioluminescent imaging-guided transcriptomic analysis. Food Chem. Toxicol. 2012, 50, 1568–1575. [Google Scholar] [CrossRef] [PubMed]
  89. Jia, R.; Cao, L.; Du, J.; Xu, P.; Jeney, G.; Yin, G. The protective effect of silymarin on the carbon tetrachloride (CCl4)-induced liver injury in common carp (Cyprinus carpio). Vitro Cell Dev. Biol. Anim. 2013, 49, 155–161. [Google Scholar] [CrossRef]
  90. Surai, P.F.; Dvorska, Y.E. Effect of mycotoxins on antioxidant status and immunity. In The Mycotoxin Blue Book; Diaz, D., Ed.; Nottingham University Press: Nottingham, UK, 2005; pp. 93–137. [Google Scholar]
  91. Sorrenti, V.; di Giacomo, C.; Acquaviva, R.; Barbagallo, I.; Bognanno, M.; Galvano, F. Toxicity of ochratoxin A and its modulation by antioxidants: A review. Toxins (Basel) 2013, 5, 1742–1766. [Google Scholar] [CrossRef]
  92. Wu, Q.H.; Wang, X.; Yang, W.; Nüssler, A.K.; Xiong, L.Y.; Kuča, K.; Dohnal, V.; Zhang, X.J.; Yuan, Z.H. Oxidative stress-mediated cytotoxicity and metabolism of T-2 toxin and deoxynivalenol in animals and humans: An update. Arch. Toxicol. 2014, 88, 1309–1326. [Google Scholar] [CrossRef] [PubMed]
  93. Essid, E.; Dernawi, Y.; Petzinger, E. Apoptosis induction by OTA and TNF-α in cultured primary rat hepatocytes and prevention by silibinin. Toxins (Basel) 2012, 4, 1139–1156. [Google Scholar] [CrossRef]
  94. Al-Anati, L.; Essid, E.; Reinehr, R.; Petzinger, E. Silibinin protects OTA-mediated TNF-alpha release from perfused rat livers and isolated rat Kupffer cells. Mol. Nutr. Food Res. 2009, 53, 460–466. [Google Scholar] [CrossRef] [PubMed]
  95. Shyamal, S.; Latha, P.G.; Suja, S.R.; Shine, V.J.; Anuja, G.I.; Sini, S.; Pradeep, S.; Shikha, P.; Rajasekharan, S. Hepatoprotective effect of three herbal extracts on aflatoxin B1-intoxicated rat liver. Singapore Med. J. 2010, 51, 326–331. [Google Scholar] [PubMed]
  96. Sozmen, M.; Devrim, A.K.; Tunca, R.; Bayezit, M.; Dag, S.; Essiz, D. Protective effects of silymarin on fumonisin B1-induced hepatotoxicity in mice. J. Vet. Sci. 2014, 15, 51–60. [Google Scholar] [CrossRef] [PubMed]
  97. Mustafa, H.N.; El Awdan, S.A.; Hegazy, G.A. Protective role of antioxidants on thioacetamide-induced acute hepatic encephalopathy: Biochemical and ultrastructural study. Tissue Cell 2013, 45, 350–362. [Google Scholar] [CrossRef] [PubMed]
  98. Fazal, Y.; Fatima, S.N.; Shahid, S.M.; Mahboob, T. Effects of curcumin on angiotensin-converting enzyme gene expression, oxidative stress and anti-oxidant status in thioacetamide-induced hepatotoxicity. J. Renin Angiotensin Aldosterone Syst. 2014. [Google Scholar] [CrossRef]
  99. Kadir, F.A.; Kassim, N.M.; Abdulla, M.A.; Kamalidehghan, B.; Ahmadipour, F.; Yehye, W.A. PASS-predicted hepatoprotective activity of Caesalpinia sappan in thioacetamide-induced liver fibrosis in rats. Sci. World J. 2014, 2014, 301879. [Google Scholar] [CrossRef]
  100. Singh, K.; Singh, N.; Chandy, A.; Manigauha, A. In vivo antioxidant and hepatoprotective activity of methanolic extracts of Daucus carota seeds in experimental animals. Asian Pac. J. Trop. Biomed. 2012, 2, 385–388. [Google Scholar] [CrossRef] [PubMed]
  101. Ali, S.O.; Darwish, H.A.; Ismail, NA. Modulatory effects of curcumin, silybin-phytosome and alpha-R-lipoic acid against thioacetamide-induced liver cirrhosis in rats. Chem. Biol. Interact. 2014, 216, 26–33. [Google Scholar] [CrossRef] [PubMed]
  102. Dos Santos, N.A.; Carvalho Rodrigues, M.A.; Martins, N.M.; dos Santos, A.C. Cisplatin-induced nephrotoxicity and targets of nephroprotection: An update. Arch. Toxicol. 2012, 86, 1233–1250. [Google Scholar] [CrossRef] [PubMed]
  103. Mansour, H.H.; Hafez, H.F.; Fahmy, N.M. Silymarin modulates Cisplatin-induced oxidative stress and hepatotoxicity in rats. J. Biochem. Mol. Biol. 2006, 39, 656–661. [Google Scholar] [CrossRef] [PubMed]
  104. El-Awady, el-S.E.; Moustafa, Y.M.; Abo-Elmatty, D.M.; Radwan, A. Cisplatin-induced cardiotoxicity: Mechanisms and cardioprotective strategies. Eur. J. Pharmacol. 2011, 650, 335–341. [Google Scholar] [CrossRef] [PubMed]
  105. Ninsontia, C.; Pongjit, K.; Chaotham, C.; Chanvorachote, P. Silymarin selectively protects human renal cells from cisplatin-induced cell death. Pharm. Biol. 2011, 49, 1082–1090. [Google Scholar] [CrossRef] [PubMed]
  106. Cho, S.I.; Lee, J.E.; Do, N.Y. Protective effect of silymarin against cisplatin-induced ototoxicity. Int. J. Pediatr. Otorhinolaryngol. 2014, 78, 474–478. [Google Scholar] [CrossRef] [PubMed]
  107. Karki, P.; Lee, E.; Aschner, M. Manganese neurotoxicity: A focus on glutamate transporters. Ann. Occup. Environ. Med. 2013, 25, 4. [Google Scholar] [CrossRef] [PubMed]
  108. Chtourou, Y.; Fetoui, H.; Sefi, M.; Trabelsi, K.; Barkallah, M.; Boudawara, T.; Kallel, H.; Zeghal, N. Silymarin, a natural antioxidant, protects cerebral cortex against manganese-induced neurotoxicity in adult rats. Biometals 2010, 23, 985–996. [Google Scholar] [CrossRef] [PubMed]
  109. Chtourou, Y.; Garoui, E.; Boudawara, T.; Zeghal, N. Therapeutic efficacy of silymarin from milk thistle in reducing manganese-induced hepatic damage and apoptosis in rats. Hum. Exp. Toxicol. 2013, 32, 70–81. [Google Scholar] [CrossRef] [PubMed]
  110. Chtourou, Y.; Fetoui, H.; Garoui, M.; Boudawara, T.; Zeghal, N. Improvement of cerebellum redox states and cholinergic functions contribute to the beneficial effects of silymarin against manganese-induced neurotoxicity. Neurochem. Res. 2012, 37, 469–479. [Google Scholar] [CrossRef] [PubMed]
  111. Chtourou, Y.; Garoui, M.; Boudawara, T.; Zeghal, N. Protective role of silymarin against manganese-induced nephrotoxicity and oxidative stress in rat. Environ. Toxicol. 2014, 29, 1147–1154. [Google Scholar] [CrossRef] [PubMed]
  112. Perumal Vijayaraman, K.; Muruganantham, S.; Subramanian, M.; Shunmugiah, K.P.; Kasi, P.D. Silymarin attenuates benzo(a)pyrene induced toxicity by mitigating ROS production, DNA damage and calcium mediated apoptosis in peripheral blood mononuclear cells (PBMC). Ecotoxicol. Environ. Saf. 2012, 86, 79–85. [Google Scholar] [CrossRef] [PubMed]
  113. Kiruthiga, P.V.; Shafreen, R.B.; Pandian, S.K.; Devi, K.P. Silymarin protection against major reactive oxygen species released by environmental toxins: Exogenous H2O2 exposure in erythrocytes. Basic Clin. Pharmacol. Toxicol. 2007, 100, 414–419. [Google Scholar] [CrossRef] [PubMed]
  114. Kiruthiga, P.V.; Pandian, S.K.; Devi, K.P. Silymarin protects PBMC against B(a)P induced toxicity by replenishing redox status and modulating glutathione metabolizing enzymes—An in vitro study. Toxicol. Appl. Pharmacol. 2010, 247, 116–128. [Google Scholar] [CrossRef] [PubMed]
  115. Kiruthiga, P.V.; Karthikeyan, K.; Archunan, G.; Karutha Pandian, S.; Pandima Devi, K. Silymarin prevents benzo(a)pyrene-induced toxicity in Wistar rats by modulating xenobiotic-metabolizing enzymes. Toxicol. Ind. Health. 2013. [Google Scholar] [CrossRef]
  116. Dirks-Naylor, A.J. The role of autophagy in doxorubicin-induced cardiotoxicity. Life Sci. 2013, 93, 913–916. [Google Scholar] [CrossRef] [PubMed]
  117. Malekinejad, H.; Janbaz-Acyabar, H.; Razi, M.; Varasteh, S. Preventive and protective effects of silymarin on doxorubicin-induced testicular damages correlate with changes in c-myc gene expression. Phytomedicine 2012, 19, 1077–1084. [Google Scholar] [CrossRef] [PubMed]
  118. El-Shitany, N.A.; El-Haggar, S.; El-Desoky, K. Silymarin prevents adriamycin-induced cardiotoxicity and nephrotoxicity in rats. Food Chem. Toxicol. 2008, 46, 2422–2428. [Google Scholar] [CrossRef] [PubMed]
  119. Patel, N.; Joseph, C.; Corcoran, G.B.; Ray, S.D. Silymarin modulates doxorubicin-induced oxidative stress, Bcl-xL and p53 expression while preventing apoptotic and necrotic cell death in the liver. Toxicol. Appl. Pharmacol. 2010, 245, 143–152. [Google Scholar] [CrossRef] [PubMed]
  120. Rašković, A.; Stilinović, N.; Kolarović, J.; Vasović, V.; Vukmirović, S.; Mikov, M. The protective effects of silymarin against doxorubicin-induced cardiotoxicity and hepatotoxicity in rats. Molecules 2011, 16, 8601–8613. [Google Scholar] [CrossRef] [PubMed]
  121. Galicia-Moreno, M.; Gutiérrez-Reyes, G. The role of oxidative stress in the development of alcoholic liver disease. Rev. Gastroenterol. Mex. 2014, 79, 135–144. [Google Scholar] [PubMed]
  122. Song, Z.; Deaciuc, I.; Song, M.; Lee, D.Y.; Liu, Y.; Ji, X.; McClain, C. Silymarin protects against acute ethanol-induced hepatotoxicity in mice. Alcohol. Clin. Exp. Res. 2006, 30, 407–413. [Google Scholar] [CrossRef] [PubMed]
  123. Zhang, B.; Buya, M.; Qin, W.; Sun, C.; Cai, H.; Xie, Q.; Xu, B.; Wu, Y. Anthocyanins from Chinese bayberry extract activate transcription factor Nrf2 in β cells and negatively regulate oxidative stress-induced autophagy. J. Agric. Food Chem. 2013, 61, 8765–8772. [Google Scholar] [CrossRef] [PubMed]
  124. Das, S.K.; Mukherjee, S. Biochemical and immunological basis of silymarin effect, a milk thistle (Silybum marianum) against ethanol-induced oxidative damage. Toxicol. Mech. Methods 2012, 22, 409–413. [Google Scholar] [CrossRef] [PubMed]
  125. Gupta, Y.K.; Sharma, M.; Chaudhary, G. Pyrogallolinduced hepatotoxicity in rats: A model to evaluate antioxidant hepatoprotective agents. Methods Find. Exp. Clin. Pharmacol. 2002, 24, 497–500. [Google Scholar] [CrossRef] [PubMed]
  126. Upadhyay, G.; Kumar, A.; Singh, M.P. Effect of silymarin on pyrogallol- and rifampicin-induced hepatotoxicity in mouse. Eur. J. Pharmacol. 2007, 565, 190–201. [Google Scholar] [CrossRef] [PubMed]
  127. Upadhyay, G.; Tiwari, M.N.; Prakash, O.; Jyoti, A.; Shanker, R.; Singh, M.P. Involvement of multiple molecular events in pyrogallol-induced hepatotoxicity and silymarin-mediated protection: Evidence from gene expression profiles. Food Chem. Toxicol. 2010, 48, 1660–1670. [Google Scholar] [CrossRef] [PubMed]
  128. Razavi-Azarkhiavi, K.; Ali-Omrani, M.; Solgi, R.; Bagheri, P.; Haji-Noormohammadi, M.; Amani, N.; Sepand, M.R. Silymarin alleviates bleomycin-induced pulmonary toxicity and lipid peroxidation in mice. Pharm. Biol. 2014, 52, 1267–1271. [Google Scholar] [CrossRef] [PubMed]
  129. Taghiabadi, E.; Imenshahidi, M.; Abnous, K.; Mosafa, F.; Sankian, M.; Memar, B.; Karimi, G. Protective Effect of Silymarin against Acrolein-Induced Cardiotoxicity in Mice. Evid. Based Complement. Alternat. Med. 2012, 2012, 352091. [Google Scholar] [CrossRef] [PubMed]
  130. Pradeep, K.; Mohan, C.V.; Gobianand, K.; Karthikeyan, S. Silymarin modulates the oxidant-antioxidant imbalance during diethylnitrosamine induced oxidative stress in rats. Eur. J. Pharmacol. 2007, 560, 110–116. [Google Scholar] [CrossRef] [PubMed]
  131. Nabavi, S.M.; Nabavi, S.F.; Moghaddam, A.H.; Setzer, W.N.; Mirzaei, M. Effect of silymarin on sodium fluoride-induced toxicity and oxidative stress in rat cardiac tissues. An. Acad. Bras. Cienc. 2012, 84, 1121–1126. [Google Scholar] [CrossRef] [PubMed]
  132. Sherif, I.O.; Al-Gayyar, M.M. Antioxidant, anti-inflammatory and hepatoprotective effects of silymarin on hepatic dysfunction induced by sodium nitrite. Eur. Cytokine Netw. 2013, 24, 114–121. [Google Scholar] [PubMed]
  133. Mateen, S.; Raina, K.; Agarwal, R. Chemopreventive and anti-cancer efficacy of silibinin against growth and progression of lung cancer. Nutr. Cancer. 2013, 65, 3–11. [Google Scholar] [CrossRef] [PubMed]
  134. Deep, G.; Agarwal, R. Targeting tumor microenvironment with silibinin: Promise and potential for a translational cancer chemopreventive strategy. Curr. Cancer Drug Targets 2013, 13, 486–499. [Google Scholar] [CrossRef] [PubMed]
  135. Ting, H.; Deep, G.; Agarwal, R. Molecular mechanisms of silibinin-mediated cancer chemoprevention with major emphasis on prostate cancer. AAPS J. 2013, 15, 707–716. [Google Scholar] [CrossRef] [PubMed]
  136. Kostek, H.; Szponar, J.; Tchórz, M.; Majewska, M.; Lewandowska-Stanek, H. Silibinin and its hepatoprotective action from the perspective of a toxicologist. Przegl. Lek. 2012, 69, 541–543. [Google Scholar] [PubMed]
  137. Féher, J.; Lengyel, G. Silymarin in the prevention and treatment of liver diseases and primary liver cancer. Curr. Pharm. Biotechnol. 2012, 13, 210–217. [Google Scholar] [CrossRef] [PubMed]
  138. Vaid, M.; Katiyar, S.K. Molecular mechanisms of inhibition of photocarcinogenesis by silymarin, a phytochemical from milk thistle (Silybum marianum L. Gaertn.) (Review). Int. J. Oncol. 2010, 36, 1053–1060. [Google Scholar] [PubMed]
  139. Kim, N.C.; Graf, T.N.; Sparacino, C.M.; Wani, M.C.; Wall, M.E. Complete isolation and characterization of silybins and isosilybins from milk thistle (Silybum. marianum). Org. Biomol. Chem. 2003, 1, 1684–1689. [Google Scholar] [CrossRef] [PubMed]
  140. Hoh, C.; Boocock, D.; Marczylo, T.; Singh, R.; Berry, D.P.; Dennison, A.R.; Hemingway, D.; Miller, A.; West, K.; Euden, S.; et al. Pilot study of oral silibinin, a putative chemopreventive agent, in colorectal cancer patients: Silibinin levels in plasma, colorectum, and liver and their pharmacodynamic consequences. Clin. Cancer Res. 2006, 12, 2944–2950. [Google Scholar] [CrossRef] [PubMed]
  141. Halliwell, B. Are polyphenols antioxidants or pro-oxidants? What do we learn from cell culture and in vivo studies? Arch. Biochem. Biophys. 2008, 476, 107–112. [Google Scholar] [CrossRef] [PubMed]
  142. Bell, J.R.; Donovan, J.L.; Wong, R.; Waterhouse, A.L.; German, J.B.; Walzem, R.L.; Kasim-Karakas, S.E. (+)-Catechin in human plasma after ingestion of a single serving of reconstituted red wine. Am. J. Clin. Nutr. 2000, 71, 103–108. [Google Scholar] [PubMed]
  143. Lapidot, T.; Harel, S.; Granit, R.; Kanner, J. Bioavailability of red wine anthocyanins as detected in human urine. J. Agric. Food Chem. 1998, 46, 4297–4302. [Google Scholar] [CrossRef]
  144. Hu, M.-L. Dietary polyphenols as antioxidants and anticancer agents: More questions than answers. Chang. Gung Med. J. 2011, 34, 449–460. [Google Scholar] [PubMed]
  145. Howard, A.N.; Constamble, B.J. The metabolism of adrenocorticotrophic hormone and ascorbic acid in the chick. Biochem. J. 1958, 69, 501–505. [Google Scholar] [PubMed]
  146. McKee, J.S.; Harrison, P.C.; Riskowski, G.L. Effects of supplemental ascorbic acid on the energy conversion of broiler chicks during heat stress and feed withdrawal. Poultry Sci. 1997, 76, 1278–1286. [Google Scholar] [CrossRef]
  147. Brenes, A.; Viveros, A.; Goni, I.; Centeno, C.; Sayago-Ayerdi, S.G.; Arija, I.; Saura-Calixto, F. Effect of grape pomace concentrate and vitamin E on digestibility of polyphenols and antioxidant activity in chickens. Poultry Sci. 2008, 87, 307–316. [Google Scholar] [CrossRef]
  148. Willemsen, H.; Swennen, Q.; Everaert, N.; Geraert, P.A.; Mercier, Y.; Stinckens, A.; Decuypere, E.; Buyse, J. Effects of dietary supplementation of methionine and its hydroxy analog dl-2-hydroxy-4-methylthio butanoic acid on growth performance, plasma hormone levels, and the redox status of broiler chickens exposed to high temperatures. Poultry Sci. 2011, 90, 2311–2320. [Google Scholar] [CrossRef]
  149. Benzie, I.F.; Szeto, Y.T.; Strain, J.J.; Tomlinson, B. Consumption of green tea causes rapid increase in plasma antioxidant power in humans. Nutr. Cancer 1999, 34, 83–87. [Google Scholar] [CrossRef] [PubMed]
  150. Halliwell, B.; Rafter, J.; Jenner, A. Health promotion by flavonoids, tocopherols, tocotrienols, and other phenols: Direct or indirect effects? Antioxidant or not? Am. J. Clin. Nutr. 2005, 81, 268S–276S. [Google Scholar] [PubMed]
  151. Lim, R.; Morwood, C.J.; Barker, G.; Lappas, M. Effect of silibinin in reducing inflammatory pathways in in vitro and in vivo models of infection-induced preterm birth. PLoS One 2014, 9, e92505. [Google Scholar] [CrossRef] [PubMed]
  152. Yan, L.J. Positive oxidative stress in aging and aging-related disease tolerance. Redox Biol. 2014, 2, 165–169. [Google Scholar] [CrossRef]
  153. Ma, Q.; He, X. Molecular basis of electrophilic and oxidative defense: Promises and perils of Nrf2. Pharmacol. Rev. 2012, 64, 1055–1081. [Google Scholar] [CrossRef] [PubMed]
  154. Majzunova, M.; Dovinova, I.; Barancik, M.; Chan, J.Y. Redox signaling in pathophysiology of hypertension. J. Biomed. Sci. 2013, 18, 69. [Google Scholar] [CrossRef]
  155. Song, P.; Zou, M.H. Redox regulation of endothelial cell fate. Cell Mol. Life Sci. 2014, 71, 3219–3239. [Google Scholar] [CrossRef] [PubMed]
  156. Kweider, N.; Huppertz, B.; Kadyrov, M.; Rath, W.; Pufe, T.; Wruck, C.J. A possible protective role of Nrf2 in preeclampsia. Ann. Anat. 2014, 196, 268–277. [Google Scholar] [CrossRef] [PubMed]
  157. Lushchak, V.I. Adaptive response to oxidative stress: Bacteria, fungi, plants and animals. Comp. Biochem. Physiol. C Toxicol. Pharmacol. 2011, 153, 175–190. [Google Scholar] [CrossRef] [PubMed]
  158. Ma, Q. Role of nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013, 53, 401–426. [Google Scholar] [CrossRef] [PubMed]
  159. Van der Wijst, M.G.; Brown, R.; Rots, M.G. Nrf2, the master redox switch: The Achilles’ heel of ovarian cancer? Biochim. Biophys. Acta 2014, 1846, 494–509. [Google Scholar] [PubMed]
  160. Tang, W.; Jiang, Y.F.; Ponnusamy, M.; Diallo, M. Role of Nrf2 in chronic liver disease. World J. Gastroenterol. 2014, 20, 13079–13087. [Google Scholar] [CrossRef] [PubMed]
  161. Howden, R. Nrf2 and cardiovascular defense. Oxid. Med. Cell Longev. 2013, 104308. [Google Scholar]
  162. Vriend, J.; Reiter, R.J. The Keap1-Nrf2-antioxidant response element pathway: A review of its regulation by melatonin and the proteasome. Mol. Cell. Endocrinol. 2015, 401, 213–220. [Google Scholar] [CrossRef] [PubMed]
  163. Keum, Y.S.; Choi, B.Y. Molecular and chemical regulation of the Keap1-Nrf2 signaling pathway. Molecules 2014, 19, 10074–10089. [Google Scholar] [CrossRef] [PubMed]
  164. Choi, B.H.; Kang, K.S.; Kwak, M.K. Effect of redox modulating NRF2 activators on chronic kidney disease. Molecules 2014, 19, 12727–12759. [Google Scholar] [CrossRef] [PubMed]
  165. Bhakkiyalakshmi, E.; Sireesh, D.; Rajaguru, P.; Paulmurugan, R.; Ramkumar, K.M. The emerging role of redox-sensitive Nrf2-Keap1 pathway in diabetes. Pharmacol. Res. 2015, 91, 104–114. [Google Scholar] [CrossRef] [PubMed]
  166. Lee, B.H.; Hsu, W.H.; Hsu, Y.W.; Pan, T.M. Dimerumic acid attenuates receptor for advanced glycation endproducts signal to inhibit inflammation and diabetes mediated by Nrf2 activation and promotes methylglyoxal metabolism into d-lactic acid. Free Radic. Biol. Med. 2013, 60, 7–16. [Google Scholar] [CrossRef] [PubMed]
  167. Zhou, S.; Sun, W.; Zhang, Z.; Zheng, Y. The role of Nrf2-mediated pathway in cardiac remodeling and heart failure. Oxid. Med. Cell Longev. 2014, 2014, 260429. [Google Scholar] [PubMed]
  168. Lee, J.M.; Calkins, M.J.; Chan, K.; Kan, Y.W.; Johnson, J.A. Identification of the NF-E2-related factor-2-dependent genes conferring protection against oxidative stress in primary cortical astrocytes using oligonucleotide microarray analysis. J. Biol. Chem. 2003, 278, 12029–12038. [Google Scholar] [CrossRef] [PubMed]
  169. Hayes, J.D.; Dinkova-Kostova, A.T. The Nrf2 regulatory network provides an interface between redox and intermediary metabolism. Trends Biochem. Sci. 2014, 39, 199–218. [Google Scholar] [CrossRef] [PubMed]
  170. Surh, Y.J. NF-κB and Nrf2 as potential chemopreventive targets of some anti-inflammatory and antioxidative phytonutrients with anti-inflammatory and antioxidative activities. Asia Pac. J. Clin. Nutr. 2008, 17, 269–272. [Google Scholar] [PubMed]
  171. Surh, Y-J.; Kundu, J.K.; Na, H.-K. Nrf2 as a master redox switch in turning on the cellular signaling involved in the induction of cytoprotective genes by some chemopreventive phytochemicals. Plant. Med. 2008, 74, 1526–1539. [Google Scholar] [CrossRef]
  172. Dayalan Naidu, S.; Kostov, R.V.; Dinkova-Kostova, A.T. Transcription factors Hsf1 and Nrf2 engage in crosstalk for cytoprotection. Trends Pharmacol. Sci. 2015, 36, 6–14. [Google Scholar] [CrossRef] [PubMed]
  173. Loboda, A.; Rojczyk-Golebiewska, E.; Bednarczyk-Cwynar, B.; Lucjusz, Z.; Jozkowicz, A.; Dulak, J. Targeting Nrf2-mediated gene transcription by triterpenoids and their derivatives. Biomol. Ther. 2012, 20, 499–505. [Google Scholar] [CrossRef]
  174. Owusu-Ansah, A.; Choi, S.H.; Petrosiute, A.; Letterio, J.J.; Huang, A.Y. Triterpenoid inducers of Nrf2 signaling as potential therapeutic agents in sickle cell disease: A review. Front. Med. 2015, 9, 46–56. [Google Scholar] [CrossRef] [PubMed]
  175. Copple, I.M.; Shelton, L.M.; Walsh, J.; Kratschmar, D.V.; Lister, A.; Odermatt, A.; Goldring, C.E.; Dinkova-Kostova, A.T.; Honda, T.; Park, B.K. Chemical tuning enhances both potency toward Nrf2 and in vitro therapeutic index of triterpenoids. Toxicol. Sci. 2014, 140, 462–469. [Google Scholar] [CrossRef] [PubMed]
  176. Velmurugan, K.; Alam, J.; McCord, J.M.; Pugazhenthi, S. Synergistic induction of heme oxygenase-1 by the components of the antioxidant supplement Protandim. Free Radic. Biol. Med. 2009, 46, 430–440. [Google Scholar] [CrossRef] [PubMed]
  177. Donovan, E.L.; McCord, J.M.; Reuland, D.J.; Miller, B.F.; Hamilton, K.L. Phytochemical activation of Nrf2 protects human coronary artery endothelial cells against an oxidative challenge. Oxid. Med. Cell Longev. 2012, 2012, 132931. [Google Scholar] [CrossRef] [PubMed]
  178. Mehrab-Mohseni, M.; Sendi, H.; Steuerwald, N.; Ghosh, S.; Schrum, L.W.; Bonkovsky, H.L. Legalon-SIL downregulates HCV core and NS5A in human hepatocytes expressing full-length HCV. World J. Gastroenterol. 2011, 17, 1694–1700. [Google Scholar] [CrossRef] [PubMed]
  179. Podder, B.; Kim, Y.S.; Zerin, T.; Song, H.Y. Antioxidant effect of silymarin on paraquat-induced human lung adenocarcinoma A549 cell line. Food Chem. Toxicol. 2012, 50, 3206–3214. [Google Scholar] [CrossRef] [PubMed]
  180. Gokila Vani, M.; Kumar, K.J.; Liao, J.W.; Chien, S.C.; Mau, J.L.; Chiang, S.S.; Lin, C.C.; Kuo, Y.H.; Wang, S.Y. Antcin C from Antrodia cinnamomea Protects Liver Cells Against Free Radical-Induced Oxidative Stress and Apoptosis In Vitro and In Vivo through Nrf2-Dependent Mechanism. Evid. Based Complement. Alternat. Med. 2013, 2013, 296082. [Google Scholar] [CrossRef] [PubMed]
  181. Kim, M.; Yang, S.G.; Kim, J.M.; Lee, J.W.; Kim, Y.S.; Lee, J.I. Silymarin suppresses hepatic stellate cell activation in a dietary rat model of non-alcoholic steatohepatitis: Analysis of isolated hepatic stellate cells. Int. J. Mol. Med. 2012, 30, 473–479. [Google Scholar] [PubMed]
  182. Wu, S.; Yue, Y.; Tian, H.; Li, Z.; Li, X.; He, W.; Ding, H. Carthamus red from Carthamus tinctorius L. exerts antioxidant and hepatoprotective effect against CCl4-induced liver damage in rats via the Nrf2 pathway. J. Ethnopharmacol. 2013, 148, 570–578. [Google Scholar] [CrossRef] [PubMed]
  183. Choi, M.K.; Han, J.M.; Kim, H.G.; Lee, J.S.; Lee, J.S.; Wang, J.H.; Son, S.W.; Park, H.J.; Son, C.G. Aqueous extract of Artemisia capillaris exerts hepatoprotective action in alcohol-pyrazole-fed rat model. J. Ethnopharmacol. 2013, 147, 662–670. [Google Scholar] [CrossRef] [PubMed]
  184. Cao, Y.W.; Jiang, Y.; Zhang, D.Y.; Wang, M.; Chen, W.S.; Su, H.; Wang, Y.T.; Wan, J.B. Protective effects of Penthorum chinense Pursh against chronic ethanol-induced liver injury in mice. J. Ethnopharmacol. 2015, 161, 92–98. [Google Scholar] [CrossRef] [PubMed]
  185. Yousefi, M.; Ghaffari, S.H.; Soltani, B.M.; Nafissi, S.; Momeny, M.; Zekri, A.; Behmanesh, M.; Alimoghaddam, K.; Ghavamzadeh, A. Therapeutic efficacy of silibinin on human neuroblastoma cells: Akt and NF-κB expressions may play an important role in silibinin-induced response. Neurochem. Res. 2012, 37, 2053–2063. [Google Scholar] [CrossRef] [PubMed]
  186. Pferschy-Wenzig, E.M.; Atanasov, A.G.; Malainer, C.; Noha, S.M.; Kunert, O.; Schuster, D.; Heiss, E.H.; Oberlies, N.H.; Wagner, H.; Bauer, R.; et al. Identification of isosilybin a from milk thistle seeds as an agonist of peroxisome proliferator-activated receptor gamma. J. Nat. Prod. 2014, 77, 842–847. [Google Scholar] [CrossRef] [PubMed]
  187. Prakash, P.; Singh, V.; Jainm, M.; Rana, M.; Khanna, V.; Barthwal, M.K.; Dikshit, M. Silymarin ameliorates fructose induced insulin resistance syndrome by reducing de novo hepatic lipogenesis in the rat. Eur. J. Pharmacol. 2014, 727, 15–28. [Google Scholar] [CrossRef] [PubMed]
  188. Lin, C.H.; Li, C.H.; Liao, P.L.; Tse, L.S.; Huang, W.K.; Cheng, H.W.; Cheng, Y.W. Silibinin inhibits VEGF secretion and age-related macular degeneration in a hypoxia-dependent manner through the PI-3 kinase/Akt/mTOR pathway. Br. J. Pharmacol. 2013, 168, 920–931. [Google Scholar] [CrossRef] [PubMed]
  189. Wang, C.; Wang, Z.; Zhang, X.; Zhang, X.; Dong, L.; Xing, Y.; Lim, Y.; Liu, Z.; Chen, L.; Qiao, H.; et al. Protection by silibinin against experimental ischemic stroke: Up-regulated pAkt, pmTOR, HIF-1α and Bcl-2, down-regulated Bax, NF-κB expression. Neurosci. Lett. 2012, 529, 45–50. [Google Scholar] [CrossRef] [PubMed]
  190. Jung, H.J.; Park, J.W.; Lee, J.S.; Lee, S.R.; Jang, B.C.; Suh, S.I.; Suh, M.H.; Baek, W.K. Silibinin inhibits expression of HIF-1alpha through suppression of protein translation in prostate cancer cells. Biochem. Biophys. Res. Commun. 2009, 390, 71–76. [Google Scholar] [CrossRef] [PubMed]
  191. Tyagi, A.; Agarwal, C.; Dwyer-Nield, L.D.; Singh, R.P.; Malkinson, A.M.; Agarwal, R. Silibinin modulates TNF-α and IFN-γ mediated signaling to regulate COX2 and iNOS expression in tumorigenic mouse lung epithelial LM2 cells. Mol. Carcinog. 2012, 51, 832–842. [Google Scholar] [CrossRef] [PubMed]
  192. Hou, Y.C.; Liou, K.T.; Chern, C.M.; Wang, Y.H.; Liao, J.F.; Chang, S.; Chou, Y.H.; Shen, Y.C. Preventive effect of silymarin in cerebral ischemia-reperfusion-induced brain injury in rats possibly through impairing NF-κB and STAT-1 activation. Phytomedicine 2010, 17, 963–973. [Google Scholar] [CrossRef] [PubMed]
  193. Agarwal, C.; Tyagi, A.; Kaur, M.; Agarwal, R. Silibinin inhibits constitutive activation of Stat3, and causes caspase activation and apoptotic death of human prostate carcinoma DU145 cells. Carcinogenesis 2007, 28, 1463–1470. [Google Scholar] [CrossRef] [PubMed]
  194. Lee-Hilz, Y.Y.; Boerboom, A.M.; Westphal, A.H.; Berkel, W.J.; Aarts, J.M.; Rietjens, I.M. Pro-oxidant activity of flavonoids induces EpRE-mediated gene expression. Chem. Res. Toxicol. 2006, 19, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  195. Erlank, H.; Elmann, A.; Kohen, R.; Kanner, J. Polyphenols activate Nrf2 in astrocytes via H2O2, semiquinones, and quinones. Free Radic. Biol. Med. 2011, 51, 2319–2327. [Google Scholar] [CrossRef] [PubMed]
  196. Tang, N.; Wu, D.; Lu, Y.; Chen, J.; Zhang, B.; Wu, W. A comparative study on the stability of silybin and that in silymarin in buffers and biological fluids. Drug Metab. Lett. 2009, 3, 115–119. [Google Scholar] [CrossRef] [PubMed]
  197. Zatloukalová, M.; Křen, V.; Gažák, R.; Kubala, M.; Trouillas, P.; Ulrichová, J.; Vacek, J. Electrochemical investigation of flavonolignans and study of their interactions with DNA in the presence of Cu(II). Bioelectrochemistry 2011, 82, 117–124. [Google Scholar] [CrossRef] [PubMed]
  198. Gharagozloo, M.; Khoshdel, Z.; Amirghofran, Z. The effect of an iron (III) chelator, silybin, on the proliferation and cell cycle of Jurkat cells: A comparison with desferrioxamine. Eur. J. Pharmacol. 2008, 589, 1–7. [Google Scholar] [CrossRef] [PubMed]
  199. Fan, S.; Qi, M.; Yu, Y.; Li, L.; Yao, G.; Tashiro, S.; Onodera, S.; Ikejima, T. P53 activation plays a crucial role in silibinin induced ROS generation via PUMA and JNK. Free Radic. Res. 2012, 46, 310–319. [Google Scholar] [CrossRef] [PubMed]
  200. Awad, H.M.; Boersma, M.G.; Boeren, S.; van Bladeren, P.J.; Vervoort, J.; Rietjens, I.M. Structure-activity study on the quinone/quinone methide chemistry of flavonoids. Chem. Res. Toxicol. 2001, 14, 398–408. [Google Scholar] [CrossRef] [PubMed]
  201. Awad, H.M.; Boersma, M.G.; Boeren, S.; van Bladeren, P.J.; Vervoort, J.; Rietjens, I.M. The regioselectivity of glutathione adduct formation with flavonoid quinone/quinone methides is pH-dependent. Chem. Res. Toxicol. 2002, 15, 343–351. [Google Scholar] [CrossRef] [PubMed]
  202. Zhang, S.; Yang, Y.; Liang, Z.; Duan, W.; Yang, J.; Yan, J.; Wang, N.; Feng, W.; Ding, M.; Nie, Y.; et al. Silybin-mediated inhibition of notch signaling exerts antitumor activity in human hepatocellular carcinoma cells. PLoS One 2013, 8, e83699. [Google Scholar] [CrossRef] [PubMed]
  203. Kim, T.H.; Woo, J.S.; Kim, Y.K.; Kim, K.H. Silibinin induces cell death through reactive oxygen species-dependent downregulation of notch-1/ERK/Akt signaling in human breast cancer cells. J. Pharmacol. Exp. Ther. 2014, 349, 268–278. [Google Scholar] [CrossRef] [PubMed]
  204. Woo, S.M.; Min, K.J.; Kim, S.; Park, J.W.; Kim, D.E.; Chun, K.S.; Kim, Y.H.; Lee, T.J.; Kim, S.H.; Choi, Y.H.; et al. Silibinin induces apoptosis of HT29 colon carcinoma cells through early growth response-1 (EGR-1)-mediated non-steroidal anti-inflammatory drug-activated gene-1 (NAG-1) up-regulation. Chem. Biol. Interact. 2014, 211, 36–43. [Google Scholar] [CrossRef] [PubMed]
  205. Tkach, K.E.; Oyler, J.E.; Altan-Bonnet, G. Cracking the NF-κB code. Sci Signal. 2014, 7, pe5. [Google Scholar] [CrossRef] [PubMed]
  206. Buelna-Chontal, M.; Zazueta, C. Redox activation of Nrf2 & NF-κB: A double end sword? Cell Signal. 2013, 25, 2548–2557. [Google Scholar] [CrossRef] [PubMed]
  207. Pedruzzi, L.M.; Stockler-Pinto, M.B.; Leite, M., Jr.; Mafra, D. Nrf2-keap1 system versus NF-κB: The good and the evil in chronic kidney disease? Biochimie 2012, 94, 2461–2466. [Google Scholar] [CrossRef] [PubMed]
  208. Pal, S.; Bhattacharjee, A.; Ali, A.; Mandal, N.C.; Mandal, S.C.; Pal, M. Chronic inflammation and cancer: Potential chemoprevention through nuclear factor kappa B and p53 mutual antagonism. J. Inflamm. (Lond.). 2014, 11, 23. [Google Scholar] [CrossRef]
  209. Hayden, M.S.; Ghosh, S. Regulation of NF-κB by TNF family cytokines. Semin. Immunol. 2014, 26, 253–266. [Google Scholar] [CrossRef] [PubMed]
  210. Gupta, S.C.; Tyagi, A.K.; Deshmukh-Taskar, P.; Hinojosa, M.; Prasad, S.; Aggarwal, B.B. Downregulation of tumor necrosis factor and other proinflammatory biomarkers by polyphenols. Arch. Biochem. Biophys. 2014, 559, 91–99. [Google Scholar] [CrossRef] [PubMed]
  211. Stevenson, D.E.; Hurst, R.D. Polyphenolic phytochemicals—Just antioxidants or much more? Cell Mol. Life. Sci. 2007, 64, 2900–2916. [Google Scholar] [CrossRef] [PubMed]
  212. Ramasamy, K.; Agarwal, R. Multitargeted therapy of cancer by silymarin. Cancer Lett. 2008, 269, 352–362. [Google Scholar] [CrossRef] [PubMed]
  213. Manna, S.K.; Mukhopadhyay, A.; Van, N.T.; Aggarwal, B.B. Silymarin suppresses TNF-induced activation of NF-κB, c-Jun N-terminal kinase, and apoptosis. J. Immunol. 1999, 163, 6800–6809. [Google Scholar] [PubMed]
  214. Tsai, M.J.; Liao, J.F.; Lin, D.Y.; Huang, M.C.; Liou, D.Y.; Yang, H.C.; Lee, H.J.; Chen, Y.T.; Chi, C.W.; Huang, W.C.; et al. Silymarin protects spinal cord and cortical cells against oxidative stress and lipopolysaccharide stimulation. Neurochem. Int. 2010, 57, 867–875. [Google Scholar] [CrossRef] [PubMed]
  215. Giorgi, V.S.; Peracoli, M.T.; Peracoli, J.C.; Witkin, S.S.; Bannwart-Castro, C.F. Silibinin modulates the NF-κB pathway and pro-inflammatory cytokine production by mononuclear cells from preeclamptic women. J. Reprod. Immunol. 2012, 95, 67–72. [Google Scholar] [CrossRef] [PubMed]
  216. Kim, B.R.; Seo, H.S.; Ku, J.M.; Kim, G.J.; Jeon, C.Y.; Park, J.H.; Jang, B.H.; Park, S.J.; Shin, Y.C.; Ko, S.G. Silibinin inhibits the production of pro-inflammatory cytokines through inhibition of NF-κB signaling pathway in HMC-1 human mast cells. Inflamm. Res. 2013, 62, 941–950. [Google Scholar] [CrossRef] [PubMed]
  217. Gharagozloo, M.; Velardi, E.; Bruscoli, S.; Agostini, M.; di Sante, M.; Donato, V.; Amirghofran, Z.; Riccardi, C. Silymarin suppress CD4+ T cell activation and proliferation: Effects on NF-κB activity and IL-2 production. Pharmacol. Res. 2010, 61, 405–409. [Google Scholar] [CrossRef] [PubMed]
  218. Oskoueian, E.; Abdullah, N.; Idrus, Z.; Ebrahimi, M.; Goh, Y.M.; Shakeri, M.; Oskoueian, A. Palm kernel cake extract exerts hepatoprotective activity in heat-induced oxidative stress in chicken hepatocytes. BMC Complement Altern. Med. 2014, 14, 368. [Google Scholar] [CrossRef] [PubMed]
  219. Wang, Q.; Zou, L.; Liu, W.; Hao, W.; Tashiro, S.; Onodera, S.; Ikejima, T. Inhibiting NF-κB activation and ROS production are involved in the mechanism of silibinin’s protection against d-galactose-induced senescence. Pharmacol. Biochem. Behav. 2011, 98, 140–149. [Google Scholar] [CrossRef] [PubMed]
  220. Salamone, F.; Galvano, F.; Marino Gammazza, A.; Paternostro, C.; Tibullo, D.; Bucchieri, F.; Mangiameli, A.; Parola, M.; Bugianesi, E.; Li Volti, G. Silibinin improves hepatic and myocardial injury in mice with nonalcoholic steatohepatitis. Dig. Liver Dis. 2012, 44, 334–342. [Google Scholar] [CrossRef] [PubMed]
  221. Liu, B.N.; Hanm, B.X.; Liu, F. Neuroprotective effect of pAkt and HIF-1 α on ischemia rats. Asian Pac. J. Trop. Med. 2014, 7, 221–225. [Google Scholar] [CrossRef] [PubMed]
  222. Aristatile, B.; Al-Assaf, A.H.; Pugalendi, K.V. Carvacrol suppresses the expression of inflammatory marker genes in d-galactosamine-hepatotoxic rats. Asian Pac. J. Trop. Med. 2013, 6, 205–211. [Google Scholar] [CrossRef] [PubMed]
  223. Salama, S.M.; Abdulla, M.A.; Alrashdi, A.S.; Hadi, A.H. Mechanism of Hepatoprotective Effect of Boesenbergia rotunda in Thioacetamide-Induced Liver Damage in Rats. Evid. Based Complement. Alternat. Med. 2013, 2013, 157456. [Google Scholar] [CrossRef] [PubMed]
  224. Abhilash, P.A.; Harikrishnan, R.; Indira, M. Ascorbic acid is superior to silymarin in the recovery of ethanol-induced inflammatory reactions in hepatocytes of guinea pigs. J. Physiol. Biochem. 2013, 69, 785–798. [Google Scholar] [CrossRef] [PubMed]
  225. Abhilash, P.A.; Harikrishnan, R.; Indira, M. Ascorbic acid suppresses endotoxemia and NF-κB signaling cascade in alcoholic liver fibrosis in guinea pigs: A mechanistic approach. Toxicol. Appl. Pharmacol. 2014, 274, 215–224. [Google Scholar] [CrossRef] [PubMed]
  226. Rattan, S.I. The nature of gerontogenes and vitagenes. Antiaging effects of repeated heat shock on human fibroblasts. Ann. N. Y. Acad. Sci. 1998, 854, 54–60. [Google Scholar] [CrossRef] [PubMed]
  227. Calabrese, V.; Boyd-Kimball, D.; Scapagnini, G.; Butterfield, D.A. Nitric oxide and cellular stress response in brain aging and neurodegenerative disorders: The role of vitagenes. Vivo 2004, 18, 245–267. [Google Scholar]
  228. Calabrese, V.; Guagliano, E.; Sapienza, M.; Panebianco, M.; Calafato, S.; Puleo, E.; Pennisi, G.; Mancuso, C.; Butterfield, D.A.; Stella, A.G. Redox regulation of cellular stress response in aging and neurodegenerative disorders: Role of vitagenes. Neurochem. Res. 2007, 32, 757–773. [Google Scholar] [CrossRef] [PubMed]
  229. Calabrese, V.; Calafato, S.; Puleo, E.; Cornelius, C.; Sapienza, M.; Morganti, P.; Mancuso, C. Redox regulation of cellular stress response by ferulic acid ethyl ester in human dermal fibroblasts: Role of vitagenes. Clin. Dermatol. 2008, 26, 358–363. [Google Scholar] [CrossRef] [PubMed]
  230. Calabrese, V.; Cornelius, C.; Mancuso, C.; Barone, E.; Calafato, S.; Bates, T.; Rizzarelli, E.; Kostova, A.T. Vitagenes, dietary antioxidants and neuroprotection in neurodegenerative diseases. Front. Biosci. 2009, 14, 376–397. [Google Scholar] [CrossRef]
  231. Calabrese, V.; Cornelius, C.; Dinkova-Kostova, A.T.; Calabrese, E.J. Vitagenes, cellular stress response, and acetylcarnitine: Relevance to hormesis. Biofactors 2009, 35, 146–160. [Google Scholar] [CrossRef] [PubMed]
  232. Calabrese, V.; Cornelius, C.; Dinkova-Kostova, A.T.; Calabrese, E.J.; Mattson, M.P. Cellular stress responses, the hormesis paradigm, and vitagenes: Novel targets for therapeutic intervention in neurodegenerative disorders. Antioxid. Redox Signal. 2010, 13, 1763–1811. [Google Scholar] [CrossRef] [PubMed]
  233. Calabrese, V.; Cornelius, C.; Dinkova-Kostova, A.T.; Iavicoli, I.; di Paola, R.; Koverech, A.; Cuzzocrea, S.; Rizzarelli, E.; Calabrese, E.J. Cellular stress responses, hermetic phytochemicals and vitagenes in aging and longevity. Biochim. Biophys. Acta 2012, 1822, 753–783. [Google Scholar]
  234. Calabrese, V.; Scapagnini, G.; Davinelli, S.; Koverech, G.; Koverech, A.; de Pasquale, C.; Salinaro, A.T.; Scuto, M.; Calabrese, E.J.; Genazzani, A.R. Sex hormonal regulation and hormesis in aging and longevity: Role of vitagenes. J. Cell Commun. Signal. 2014, 8, 369–384. [Google Scholar] [CrossRef] [PubMed]
  235. Cornelius, C.; Perrotta, R.; Graziano, A.; Calabrese, E.J.; Calabrese, V. Stress responses, vitagenes and hormesis as critical determinants in aging and longevity: Mitochondria as a “chi”. Immun. Ageing 2013, 10, 15. [Google Scholar] [CrossRef] [PubMed]
  236. Cornelius, C.; Graziano, A.; Calabrese, E.J.; Calabrese, V. Hormesis and vitagenes in aging and longevity: Mitochondrial control and hormonal regulation. Horm. Mol. Biol. Clin. Investig. 2013, 16, 73–89. [Google Scholar] [PubMed]
  237. Cornelius, C.; Koverech, G.; Crupi, R.; di Paola, R.; Koverech, A.; Lodato, F.; Scuto, M.; Salinaro, A.T.; Cuzzocrea, S.; Calabrese, E.J.; et al. Osteoporosis and Alzheimer pathology: Role of cellular stress response and hormetic redox signaling in aging and bone remodeling. Front. Pharmacol. 2014, 5, 120. [Google Scholar] [CrossRef] [PubMed]
  238. Trovato Salinaro, A.; Cornelius, C.; Koverech, G.; Koverech, A.; Scuto, M.; Lodato, F.; Fronte, V.; Muccilli, V.; Reibaldi, M.; Longo, A.; et al. Cellular stress response, redox status, and vitagenes in glaucoma: A systemic oxidant disorder linked to Alzheimer’s disease. Front. Pharmacol. 2014, 5, 129. [Google Scholar] [CrossRef] [PubMed]
  239. Cerný, D.; Canová, N.K.; Martínek, J.; Horínek, A.; Kmonícková, E.; Zídek, Z.; Farghali, H. Effects of resveratrol pretreatment on tert-butylhydroperoxide induced hepatocyte toxicity in immobilized perifused hepatocytes: Involvement of inducible nitric oxide synthase and hemoxygenase-1. Nitric Oxide 2009, 20, 1–8. [Google Scholar] [CrossRef] [PubMed]
  240. Venditti, C.C.; Smith, G.N. Involvement of the heme oxygenase system in the development of preeclampsia and as a possible therapeutic target. Womens. Health (Lond. Engl.) 2014, 10, 623–643. [Google Scholar] [CrossRef]
  241. Naito, Y.; Takagi, T.; Higashimura, Y. Heme oxygenase-1 and anti-inflammatory M2 macrophages. Arch. Biochem. Biophys. 2014, 564, 83–88. [Google Scholar] [CrossRef] [PubMed]
  242. Ambegaokar, S.S.; Kolson, D.L. Heme oxygenase-1 dysregulation in the brain: Implications for HIV-associated neurocognitive disorders. Curr. HIV Res. 2014, 12, 174–188. [Google Scholar] [CrossRef] [PubMed]
  243. Bongiovanni, G.A.; Soria, E.A.; Eynard, A.R. Effects of the plant flavonoids silymarin and quercetin on arsenite-induced oxidative stress in CHO-K1 cells. Food Chem. Toxicol. 2007, 45, 971–976. [Google Scholar] [CrossRef] [PubMed]
  244. Demir, M.; Amanvermez, R.; Kamalı Polat, A.; Karabıçak, I.; Cınar, H.; Kesicioğlu, T.; Polat, C. The effect of silymarin on mesenteric ischemia-reperfusion injury. Med. Princ. Pract. 2014, 23, 140–144. [Google Scholar] [PubMed]
  245. Zhao, H.; Brandt, G.E.; Galam, L.; Matts, R.L.; Blagg, B.S. Identification and initial SAR of silybin: An Hsp90 inhibitor. Bioorg. Med. Chem. Lett. 2011, 21, 2659–2664. [Google Scholar] [CrossRef] [PubMed]
  246. Jiang, Y.Y.; Wang, H.J.; Wang, J.; Tashiro, S.; Onodera, S.; Ikejima, T. The protective effect of silibinin against mitomycin C-induced intrinsic apoptosis in human melanoma A375-S2 cells. J. Pharmacol. Sci. 2009, 111, 137–146. [Google Scholar] [CrossRef] [PubMed]
  247. Zhou, B.; Wu, L.J.; Li, L.H.; Tashiro, S.; Onodera, S.; Uchiumi, F.; Ikejima, T. Silibinin protects against isoproterenol-induced rat cardiac myocyte injury through mitochondrial pathway after up-regulation of SIRT1. J. Pharmacol.Sci. 2006, 102, 387–395. [Google Scholar] [CrossRef] [PubMed]
  248. Wang, Q.; Liu, M.; Liu, W.W.; Hao, W.B.; Tashiro, S.; Onodera, S.; Ikejima, T. In vivo recovery effect of silibinin treatment on streptozotocin-induced diabetic mice is associated with the modulations of Sirt-1 expression and autophagy in pancreatic β-cell. J. Asian Nat. Prod. Res. 2012, 14, 413–423. [Google Scholar] [CrossRef] [PubMed]
  249. Li, L.H.; Wu, L.J.; Tashiro, S.I.; Onodera, S.; Uchiumi, F.; Ikejima, T. Activation of the SIRT1 pathway and modulation of the cell cycle were involved in silymarin’s protection against UV-induced A375-S2 cell apoptosis. J. Asian Nat. Prod. Res. 2007, 9, 245–252. [Google Scholar] [CrossRef] [PubMed]
  250. Wang, H.J.; Tashiro, S.; Onodera, S.; Ikejima, T. Inhibition of insulin-like growth factor 1 receptor signaling enhanced silibinin-induced activation of death receptor and mitochondrial apoptotic pathways in human breast cancer MCF-7 cells. J. Pharmacol. Sci. 2008, 107, 260–269. [Google Scholar] [CrossRef] [PubMed]
  251. Wu, Y.T.; Wu, S.B.; Wei, Y.H. Roles of sirtuins in the regulation of antioxidant defense and bioenergetic function of mitochondria under oxidative stress. Free Radic. Res. 2014, 48, 1070–1084. [Google Scholar] [CrossRef] [PubMed]
  252. Wu, Y.T.; Wu, S.B.; Wei, Y.H. Metabolic reprogramming of human cells in response to oxidative stress: Implications in the pathophysiology and therapy of mitochondrial diseases. Curr. Pharm. Des. 2014, 20, 5510–5526. [Google Scholar] [CrossRef] [PubMed]
  253. Horio, Y.; Hayashi, T.; Kuno, A.; Kunimoto, R. Cellular and molecular effects of sirtuins in health and disease. Clin. Sci. (Lond.) 2011, 121, 191–203. [Google Scholar] [CrossRef]
  254. Jayasena, T.; Poljak, A.; Smythe, G.; Braidy, N.; Münch, G.; Sachdev, P. The role of polyphenols in the modulation of sirtuins and other pathways involved in Alzheimer’s disease. Ageing Res. Rev. 2013, 12, 867–883. [Google Scholar] [CrossRef] [PubMed]
  255. Radak, Z.; Koltai, E.; Taylor, A.W.; Higuchi, M.; Kumagai, S.; Ohno, H.; Goto, S.; Boldogh, I. Redox-regulating sirtuins in aging, caloric restriction, and exercise. Free Radic. Biol. Med. 2013, 58, 87–97. [Google Scholar] [CrossRef] [PubMed]
  256. Lu, J.; Holmgren, A. The thioredoxin antioxidant system. Free Radic. Biol. Med. 2014, 66, 75–87. [Google Scholar] [CrossRef] [PubMed]
  257. Penney, R.B.; Roy, D. Thioredoxin-mediated redox regulation of resistance to endocrine therapy in breast cancer. Biochim. Biophys. Acta 2013, 1836, 60–79. [Google Scholar] [PubMed]
  258. Surai, K.P.; Surai, P.F.; Speake, B.K.; Sparks, N.H.C. Antioxidant-prooxidant balance in the intestine: Food for thought. 2. Antioxidants. Curr. Top. Nutraceutical. Res. 2004, 2, 27–46. [Google Scholar]
  259. Surai, K.P.; Surai, P.F.; Speake, B.K.; Sparks, N.H.C. Antioxidant-prooxidant balance in the intestine: Food for thought. 1. Prooxidants. Nutr. Genomics Funct. Foods 2003, 1, 51–70. [Google Scholar]
  260. Dolara, P.; Luceri, C.; de Filippo, C.; Femia, A.P.; Giovannelli, L.; Caderni, G.; Cecchini, C.; Silvi, S.; Orpianesi, C.; Cresci, A. Red wine polyphenols influence carcinogenesis, intestinal microflora, oxidative damage and gene expression profiles of colonic mucosa in F344 rats. Mutat. Res. 2005, 591, 237–246. [Google Scholar] [CrossRef] [PubMed]
  261. Kanner, J.; Lapidot, T. The stomach as a bioreactor: Dietary lipid peroxidation in the gastric fluid and the effects of plant-derived antioxidants. Free Radic. Biol. Med. 2001, 31, 1388–1395. [Google Scholar] [CrossRef] [PubMed]
  262. Giovannelli, L.; Testa, G.; de Filippo, C.; Cheynier, V.; Clifford, M.N.; Dolara, P. Effect of complex polyphenols and tannins from red wine on DNA oxidative damage of rat colon mucosa in vivo. Eur. J. Nutr. 2000, 39, 207–212. [Google Scholar] [CrossRef] [PubMed]
  263. Gorelik, S.; Lapidot, T.; Shaham, I.; Granit, R.; Ligumsky, M.; Kohen, R.; Kanner, J. Lipid peroxidation and coupled vitamin oxidation in simulated and human gastric fluid inhibited by dietary polyphenols: Health implications. J. Agric. Food Chem. 2005, 53, 3397–3402. [Google Scholar] [CrossRef] [PubMed]
  264. Dai, F.; Chen, W.F.; Zhou, B. Antioxidant synergism of green tea polyphenols with alpha-tocopherol and l-ascorbic acid in SDS micelles. Biochimie 2008, 90, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  265. Gorelik, S.; Ligumsky, M.; Kohen, R.; Kanner, J. The stomach as a “bioreactor”: When red meat meets red wine. J. Agric. Food Chem. 2008, 56, 5002–5007. [Google Scholar] [CrossRef] [PubMed]
  266. Gorelik, S.; Ligumsky, M.; Kohen, R.; Kanner, J. A novel function of red wine polyphenols in humans: Prevention of absorption of cytotoxic lipid peroxidation products. FASEB J. 2008, 22, 41–46. [Google Scholar] [CrossRef] [PubMed]
  267. Kanner, J.; Gorelik, S.; Roman, S.; Kohen, R. Protection by polyphenols of postprandial human plasma and low-density lipoprotein modification: The stomach as a bioreactor. J. Agric. Food Chem. 2012, 60, 8790–8796. [Google Scholar] [CrossRef] [PubMed]
  268. Circu, M.L.; Aw, T.Y. Intestinal redox biology and oxidative stress. Semin. Cell Dev. Biol. 2012, 23, 729–737. [Google Scholar] [CrossRef] [PubMed]
  269. Martínez, J.A.; Etxeberría, U.; Galar, A.; Milagro, F.I. Role of dietary polyphenols and inflammatory processes on disease progression mediated by the gut microbiota. Rejuvenation Res. 2013, 16, 435–437. [Google Scholar] [CrossRef] [PubMed]
  270. Biasi, F.; Deiana, M.; Guina, T.; Gamba, P.; Leonarduzzi, G.; Poli, G. Wine consumption and intestinal redox homeostasis. Redox Biol. 2014, 2, 795–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Biasi, F.; Astegiano, M.; Maina, M.; Leonarduzzi, G.; Poli, G. Polyphenol supplementation as a complementary medicinal approach to treating inflammatory bowel disease. Curr. Med. Chem. 2011, 18, 4851–4865. [Google Scholar] [CrossRef] [PubMed]
  272. Gessner, D.K.; Fiesel, A.; Most, E.; Dinges, J.; Wen, G.; Ringseis, R.; Eder, K. Supplementation of a grape seed and grape marc meal extract decreases activities of the oxidative stress-responsive transcription factors NF-κB and Nrf2 in the duodenal mucosa of pigs. Acta Vet. Scand. 2013, 55, 18. [Google Scholar] [CrossRef] [PubMed]
  273. Yang, G.; Wangm, H.; Kang, Y.; Zhu, M.J. Grape seed extract improves epithelial structure and suppresses inflammation in ileum of IL-10-deficient mice. Food Funct. 2014, 5, 2558–2563. [Google Scholar] [CrossRef] [PubMed]
  274. Goodrich, K.M.; Fundaro, G.; Griffin, L.E.; Grant, A.; Hulver, M.W.; Ponder, M.A.; Neilson, A.P. Chronic administration of dietary grape seed extract increases colonic expression of gut tight junction protein occludin and reduces fecal calprotectin: A secondary analysis of healthy Wistar Furth rats. Nutr. Res. 2012, 32, 787–794. [Google Scholar] [CrossRef] [PubMed]
  275. Rodríguez-Ramiro, I.; Ramos, S.; Bravo, L.; Goya, L.; Martín, M.Á. Procyanidin B2 and a cocoa polyphenolic extract inhibit acrylamide-induced apoptosis in human Caco-2 cells by preventing oxidative stress and activation of JNK pathway. J. Nutr. Biochem. 2011, 22, 1186–1194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Rodríguez-Ramiro, I.; Martín, M.A.; Ramos, S.; Bravo, L.; Goya, L. Comparative effects of dietary flavanols on antioxidant defences and their response to oxidant-induced stress on Caco2 cells. Eur. J. Nutr. 2011, 50, 313–322. [Google Scholar] [CrossRef] [PubMed]
  277. Köck, K.; Xie, Y.; Hawke, R.L.; Oberlies, N.H.; Brouwer, K.L. Interaction of silymarin flavonolignans with organic anion-transporting polypeptides. Drug Metab. Dispos. 2013, 41, 958–965. [Google Scholar] [CrossRef] [PubMed]
  278. Shin, J.H.; Lee, C.W.; Oh, S.J.; Yun, J.; Lee, K.; Park, S.K.; Kim, H.M.; Han, S.B.; Kim, Y.; Kim, H.C.; et al. Protective effect of silymarin against ethanol-induced gastritis in rats: Role of sulfhydryls, nitric oxide and gastric sensory afferents. Food Chem. Toxicol. 2013, 55, 353–357. [Google Scholar] [CrossRef] [PubMed]
  279. Esmaily, H.; Hosseini-Tabatabaei, A.; Rahimian, R.; Khorasani, R.; Baeeri, M.; Barazesh-Morgani, A.; Yasa, N.; Khademi, Y.; Abdollahi, M. On the benefits of silymarin in murine colitis by improving balance of destructive cytokines and reduction of toxic stress in the bowel cells. Central Eur. Kournal. Biol. 2009, 4, 204–213. [Google Scholar] [CrossRef]
  280. Esmaily, H.; Vaziri-Bami, A.; Miroliaee, A.E.; Baeeri, M.; Abdollahi, M. The correlation between NF-κB inhibition and disease activity by coadministration of silibinin and ursodeoxycholic acid in experimental colitis. Fundam. Clin. Pharmacol. 2011, 25, 723–733. [Google Scholar] [CrossRef] [PubMed]
  281. Moco, S.; Martin, F.P.; Rezzi, S. Metabolomics view on gut microbiome modulation by polyphenol-rich foods. J. Proteome. Res. 2012, 11, 4781–4790. [Google Scholar] [CrossRef] [PubMed]
  282. Pozuelo, M.J.; Agis-Torres, A.; Hervert-Hernández, D.; Elvira López-Oliva, M.; Muñoz-Martínez, E.; Rotger, R.; Goñi, I. Grape antioxidant dietary fiber stimulates Lactobacillus growth in rat cecum. J. Food Sci. 2012, 77, H59–H62. [Google Scholar] [CrossRef] [PubMed]
  283. Parkar, S.G.; Trower, T.M.; Stevenson, D.E. Fecal microbial metabolism of polyphenols and its effects on human gut microbiota. Anaerobe 2013, 23, 12–19. [Google Scholar] [CrossRef] [PubMed]
  284. Bolca, S.; van de Wiele, T.; Possemiers, S. Gut metabotypes govern health effects of dietary polyphenols. Curr. Opin. Biotechnol. 2013, 24, 220–225. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Surai, P.F. Silymarin as a Natural Antioxidant: An Overview of the Current Evidence and Perspectives. Antioxidants 2015, 4, 204-247. https://doi.org/10.3390/antiox4010204

AMA Style

Surai PF. Silymarin as a Natural Antioxidant: An Overview of the Current Evidence and Perspectives. Antioxidants. 2015; 4(1):204-247. https://doi.org/10.3390/antiox4010204

Chicago/Turabian Style

Surai, Peter F. 2015. "Silymarin as a Natural Antioxidant: An Overview of the Current Evidence and Perspectives" Antioxidants 4, no. 1: 204-247. https://doi.org/10.3390/antiox4010204

Article Metrics

Back to TopTop